You are on page 1of 35

An Integrated Index: Engrams, Place

Cells, and Hippocampal Memory Travis


D. Goode & Kazumasa Z. Tanaka &
Amar Sahay & Thomas J. Mchugh
Visit to download the full and correct content document:
https://ebookmass.com/product/an-integrated-index-engrams-place-cells-and-hippoca
mpal-memory-travis-d-goode-kazumasa-z-tanaka-amar-sahay-thomas-j-mchugh/
Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll

Perspective
An Integrated Index: Engrams,
Place Cells, and Hippocampal Memory
Travis D. Goode,1,2,3,4,7 Kazumasa Z. Tanaka,5,7 Amar Sahay,1,2,3,4,* and Thomas J. McHugh6,*
1Center for Regenerative Medicine, Massachusetts General Hospital, Boston, MA 02114, USA
2Harvard Stem Cell Institute, Cambridge, MA 02138, USA
3Department of Psychiatry, Massachusetts General Hospital, Harvard Medical School, Boston, MA 02114 USA
4Broad Institute of Harvard and MIT, Cambridge, MA 02142, USA
5Memory Research Unit, Okinawa Institute of Science and Technology Graduate University, Onna-son, Kunigami-gun, Okinawa, Japan
6Laboratory for Circuit and Behavioral Physiology, RIKEN Center for Brain Science, Wakoshi, Saitama, Japan
7These authors contributed equally

*Correspondence: asahay@mgh.harvard.edu (A.S.), thomas.mchugh@riken.jp (T.J.M.)


https://doi.org/10.1016/j.neuron.2020.07.011

SUMMARY

The hippocampus and its extended network contribute to encoding and recall of episodic experiences.
Drawing from recent anatomical, physiological, and behavioral studies, we propose that hippocampal en-
grams function as indices to mediate memory recall. We broaden this idea to discuss potential relationships
between engrams and hippocampal place cells, as well as the molecular, cellular, physiological, and circuit
determinants of engrams that permit flexible routing of information to intra- and extrahippocampal circuits for
reinstatement, a feature critical to memory indexing. Incorporating indexing into frameworks of memory func-
tion opens new avenues of study and even therapies for hippocampal dysfunction.

Introduction Space and Memory


A few keywords typed into the Google search bar will, more often The discovery of place cells, neurons in the hippocampus that
than not, immediately lead to the exact piece of information we have receptive fields tuned to discrete locations within a context
are seeking. While the details of how this magic happens are pro- (O’Keefe and Dostrovsky, 1971), revolutionized the experimental
prietary, the general idea is transparent; Google has managed to approach to studying hippocampal function (Moser et al., 2017).
index vast swaths of the internet and uses our search terms to One of the earliest and most influential theories to emerge linking
quickly point to the most appropriate information (https://www. place cell activity and episodic memory was the cognitive map
google.com/search/howsearchworks). The system is surpris- theory (O’Keefe and Nadel, 1978), positing that the primary
ingly flexible, using history, context, or location to hone results; role of the hippocampus is to provide a spatial framework that
completing or anticipating partial bits of information; and finding permits the location and association of the items and events
and separating similar items by detecting small differences. that constitute a given experience. While the authors suggested
These properties, which underlie both its efficiency and popu- that this cognitive map may not be limited to physical space and
larity, echo the abilities of the memory systems operating in could be applied to map episodic experience more broadly,
our own brains, particularly the episodic memory circuits depen- place cells and their properties have proved a useful substrate
dent on the hippocampus (Squire et al., 2004; Tulving, 2002). The to examine and test these ideas. For example, as different en-
hippocampus is crucial for the encoding of memory, it is adept at sembles of neurons are recruited to represent different spatial
integrating and interpreting contextual cues to drive recall, and it experiences, the hippocampus could continuously provide a
is efficient at both discrimination and association (Maren et al., new underlying scaffolding across space (and time) that would
2013). Thus, much like how Google works as an index of informa- allow memories to remain both related and distinct. Subsequent
tion, one parsimonious explanation for hippocampal function is retrieval could then be triggered by a reinstatement of the original
that it functions as an index of memories (Guo et al., 2018; Miller spatial map triggered by the cues that define a given context (Wi-
and Sahay, 2019; Tanaka, 2020; Tanaka and McHugh, 2018; kenheiser and Redish, 2015). Over time, other theoretical models
Tanaka et al., 2018; Tonegawa et al., 2018). This is not a new have built on and expanded these ideas, and, as noted in other
idea (McClelland et al., 1995; Teyler and DiScenna, 1985, sections below, have proposed anatomical substrates for hippo-
1986; Teyler and Rudy, 2007); however, recent work has begun campal functions that include novelty detection (Lisman and Ot-
to lend direct experimental evidence to this theory and edifying makhova, 2001; Vinogradova, 2001), rapid encoding, and
putative, underlying circuit mechanisms. Here, we will explore pattern completion and separation (Kesner and Rolls, 2015;
and examine these findings in depth and discuss possible index- Knierim and Neunuebel, 2016; Nadel and Moscovitch, 1997),
ing mechanisms, as well as how these ideas could shape a better as well as their relation to spatial coding. Moreover, frameworks
understanding of memory processes in both the healthy and that encompass the place cell data but are not tied to a specific
diseased brain. spatial function of the hippocampus have also been described.

Neuron 107, September 9, 2020 ª 2020 Elsevier Inc. 1


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

These posit that place cells may reflect a broader functionality rience. While many engram studies have focused on measures
related to general-purpose sequence generation (Buzsáki and of conditioned fear, these and other studies report hippocampal
Tingley, 2018) or relational memory (Cohen and Eichenbaum, engram-driven behavior for a variety of context-specific behav-
1993; Eichenbaum et al., 1994), allowing the extension of both iors, including place (e.g., Ramirez et al., 2013) or social avoid-
hippocampal memory and physiology (Aronov et al., 2017; Mac- ance (e.g., Zhang et al., 2019), as well as appetitive conditioning
Donald et al., 2011; Pastalkova et al., 2008) beyond the domain and place preference (e.g., Redondo et al., 2014).
of physical space. These findings raise several key questions. First, how can a
Hand in hand with the growing characterization of place cell small number of experience-tagged hippocampal engram cells,
physiology, there developed a deeper understanding of the much fewer in fraction than that of active place cells (e.g., Tanaka
anatomical, behavioral, and computational properties of the hip- et al., 2018), encode a complex behavioral experience? Indeed,
pocampal circuit. This resulted in specific mnemonic functions activation of a very small percentage (2%–3%) of DG granule
being linked to the anatomic and physiological properties of cells labeled during learning can reproduce context-appropriate
discrete subregions of the structure (Fanselow and Dong, 2010; behaviors (for examples, see Liu et al., 2012). Further, activation
Nadel et al., 2013; Strange et al., 2014). In this framework, the in the DG could harness the pattern completion abilities of the
classic model of sequential processing along the trisynaptic downstream CA3 network to amplify their activity via attractor
loop has the large number and sparse activity of granule cells in dynamics (Colgin et al., 2010; Knierim and Neunuebel, 2016)
the dentate gyrus (DG) providing orthogonalization of similar and lead to a robust brain-wide reinstatement of a memory-
cortical inputs leading to pattern separation (Hainmueller and Bar- related ensemble. However, reactivation of a subset of CA1
tos, 2020; Leal and Yassa, 2018; McHugh et al., 2007), the DG neurons, which lack recurrent connectivity, can also trigger
providing input to the recurrent CA3 network to facilitate autoas- behavioral reinstatement and presumably memory recall (e.g.,
sociation and pattern completion (Cayco-Gajic and Silver, 2019; Ryan et al., 2015). It is plausible that downstream activation of
Kesner and Rolls, 2015; Knierim and Neunuebel, 2016; McHugh the entorhinal cortex (EC) and/or re-entrant excitation of the
et al., 2007; Nakazawa et al., 2002), and finally, CA1 broadcasting DG and CA3 add these features, thereby functioning like a recur-
the results back to the cortex (Soltesz and Losonczy, 2018; Valero rent network, although experimental evidence supporting this
and de la Prida, 2018). This framework has served as the back- interpretation is lacking.
bone of relating place cell activity to memory processing, insofar Additionally, 50 years of hippocampal physiology in rodents
as place cells may coordinate the pattern separation, completion, has revealed that place cell activity is exquisitely structured
and reinstatement properties noted above. across not only space but also time (Howard and Eichenbaum,
Building on this framework, advances in genetic approaches 2015). During exploration, the dominant theta oscillation in the
have led to activity-dependent memory tagging systems in hippocampal local field potential (LFP) organizes ensembles of
rodent models, which allow for the examination and artificial re- place cells with spatially adjacent receptive fields into se-
activation or inhibition of distinct memory traces (Josselyn and quences, expressed on the timescale of a single theta cycle of
Tonegawa, 2020). These traces fulfill the properties of the mem- 125 ms (Burgess and O’Keefe, 2011). These sequences can
ory engram, a moniker for the physical basis of memory first pro- be re-expressed during sharp-wave ripples (SWRs) that occur
posed by zoologist and biologist Richard Semon (Schacter et al., during pauses in movement on an even shorter timescale, com-
1978; Semon, 1921), in that they can be viewed as the physical pressed into fast events lasting only 10s of milliseconds (Foster,
instantiation of an experience registered in enduring changes 2017). This precise temporal arrangement of activity has made
in synaptic connectivity and physiology of an ensemble of the gap between place-cell- and engram-based memory studies
neurons. Tagging systems employed include the tetracycline- difficult to bridge, as the latter have demonstrated that
regulated transcriptional activation system, in which time-locked simultaneous optogenetic activation of ensembles of neurons
expression of actuators such as opsins or chemogenetic recep- in temporal and spatial patterns that are not observed under nat-
tors are induced via activity-dependent promoters (e.g., c-Fos- ural physiological conditions are sufficient to evoke behaviors
or Arc-expressing; Liu et al., 2012) or immediate early gene mimicking memory recall. One can interpret this gap in the tem-
(IEG)-binding elements (Sun et al., 2020), as well as the Cre- poral dynamics between optically induced behavioral reinstate-
ERT (estrogen receptor T2) transcription system (targeted ment and place cell activity as reflecting the dispensability of
recombination in active populations [TRAP] mice), which utilizes these temporal pattern for behaviors driven by contextual recog-
a tamoxifen-sensitive modified estrogen receptor to drive expe- nition, or perhaps this disconnect could simply be due to tech-
rience-driven expression of Cre-dependent constructs in acti- nical limitations in the place cell recording, as the retrieval of a
vated cells (Guenthner et al., 2013). Such methods have allowed hippocampal-dependent memory can occur very rapidly and in
for the artificial triggering of memory-related behavior even in the absence of the exploration needed to drive extensive place
contexts where no such behavior would be expected. While cell activity, precluding a robust sampling of activity. For
these methods are not without their caveats (discussed further instance, when rodents receive a footshock immediately after
in sections below), engram-labeling studies have shown that op- placement in a previously exposed chamber, context explora-
togenetic or chemogenetic stimulation or inhibition of excitatory tion may be minimal, yet animals successfully retrieve the
neurons in the DG (e.g., Guo et al., 2018; Lacagnina et al., 2019; contextual memory and associate it with shock, resulting in
Liu et al., 2012), CA3 (e.g., Denny et al., 2014), or CA1 (e.g., context-dependent behavior during subsequent memory tests
Ghandour et al., 2019; Ryan et al., 2015; Tanaka et al., 2014) re- (Wiltgen et al., 2001). One possibility is that reinstatement of
instates or impedes (respectively) behavioral recall of that expe- even a single place field is sufficient for memory reinstatement.

2 Neuron 107, September 9, 2020


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

across several weeks, it becomes more difficult to draw a direct


connection between memory recall and a completely stable
spatial representation. It is possible that a small fraction of stable
place cells can serve as a partial cue to reinstate the full repre-
sentation of memory through a process of pattern completion
in downstream regions; however, this view is challenged by a
recent study examining the physiological nature of engram cells
in the hippocampus, which is discussed below (Tanaka et al.,
2018). Thus, while place cell studies have provided insight into
the anatomical organization and potential memory mechanisms
of the hippocampus, we, like many others, struggle to reconcile
these potential spatial coding properties with the core role of the
system as a memory storage device (Tanaka, 2020). Further, it is
important, both through hypothesis and experiments, to attempt
to identify the rules of transformation that allow simultaneous
activation of hippocampal engrams to generate appropriate pat-
terns of downstream activation and behavior (Lisman et al.,
2017). Perhaps then, we should reconsider what the activity of
hippocampal neurons during memory formation and recall truly
represents and how place cells that have driven much of the
thinking in the field for the last 50 years can inform us about
the hippocampus as a memory system (Figure 1).

Instantiating the Hippocampal Index


The hippocampal memory indexing theory posits that the hippo-
campus does not ‘‘contain’’ the episodic memory itself; rather, it
generates a code or ‘‘index’’ that binds neuronal activity patterns
underlying an experiential event, which is stored across distrib-
uted neocortical (and potentially subcortical) modules (Teyler
and DiScenna, 1985, 1986; Teyler and Rudy, 2007). In other
words, the hippocampus encodes a linked representation of
brain activity at the time of an experience or episode, which
can subserve subsequent recall via activation of that hippocam-
pal representation. What is presumed to make these patterns of
activity unique from other experience-induced patterns in the
brain, such as ensemble activity in the sensory cortex activated
by a stimulus, is their conjunctive and associational nature and
the ability of the hippocampal ensemble to reinstate the original
spatial and temporal patterns of cortical/subcortical activity of an
experience (McClelland et al., 1995; Teyler and Rudy, 2007).
Figure 1. Comparing and Contrasting Hippocampal Place Cells and
Engram-Tagged (Immediate Early-Gene-Expressing) Neurons Important to note is that the indexing theory is not mutually
Place cell activity has precise temporal structure during both exploration and exclusive to the cognitive map theory. Instead, it simply remains
rest, whereas engram-tagged neurons are simultaneously and experimentally agnostic to what, if anything, the hippocampal neurons involved
reactivated. Place cell density appears moderate, and this density of active
place cells is relatively stable in any given context. Conversely, engram-tagged
in memory indexing must represent in terms of behaviorally rele-
cells in the hippocampus are sparse, with familiar contexts exhibiting low vant information; spatial coding would be acceptable if these
levels of tagged expression. Engram cells exhibit considerably less stability in neurons had properties consistent with that of an index, as sum-
their context-dependent reinstatement over time as compared to place cells,
marized and presented in Figure 2. In the following and subse-
although both are highly unstable with time. Remote time points for re-
activation of experience-dependent engram cells remain unknown. While quent sections, we elaborate on these features and discuss
there are overlapping behavioral correlates of place and engram cell activity, how the brain’s circuit architecture supports a view of hippocam-
place cell research has led the field in its correlation to behavior. pal function through the lens of indexing.
Engrams as Indices
However, long-term monitoring of the stability of place cell Numerous studies have now shown that photoactivation of a
representations across repeated visits on the timescale of sparse hippocampal engram drives IEG activity in select down-
weeks, now possible due to advances in in vivo imaging ap- stream brain regions thought to be involved in the original
proaches in mice, suggests that there exists a hitherto unappre- learning (e.g., Ramirez et al., 2013, 2015; Roy et al., 2017).
ciated high degree of instability in the spatial representation of a Such observations, together with the reinstatement of behavior
familiar environment (Ziv et al., 2013; but, also see (Gonzalez following optogenetic engram stimulation, support the idea
et al., 2019). If only a fraction (15%) of place cells show stability that the hippocampus is capable of indexing and triggering

Neuron 107, September 9, 2020 3


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective
Figure 2. Key Features of Hippocampal
Memory Indexing Theory

learning increased mossy fiber synaptic


contacts of tagged engram-bearing den-
tate granule cells (DGCs) with parvalbu-
min-positive (PV+) stratum lucidum inhib-
itory neurons (SLINs) to a significantly
greater extent than a random population
of DGCs. This engram-dependent
recruitment of PV+ SLINs returned to
pre-learning levels with time-dependent
memory generalization. By genetically
enhancing the coupling of engram-
bearing DGCs with PV+ SLINs, the au-
thors increased feed-forward inhibition
in DG-CA3 circuitry and stabilized the
hippocampal engram. Critically, this was
shown to confer optogenetic behavioral
reinstatement and context-specific reac-
tivation of a distributed fear memory trace
in hippocampal-cortical-subcortical net-
works at remote time points. Collectively,
these findings mirror natural recall, inso-
far as contextual fear memory retrieval
in the original learning context is associ-
ated with the specific reactivation of
learning-dependent tagging in the hippo-
campus and some extrahippocampal tar-
gets. However, formal demonstration for
how hippocampus may in fact coordinate
extrahippocampal reinstatement in an
experience-dependent manner is absent.
In light of understanding hippocampal
engram functions through the lens of in-
memory recall by reinstating learning-dependent activity in dexing, it is important to emphasize that although behavioral
memory-related extrahippocampal brain systems. However, reinstatement does not equate to neural reinstatement, the
increased IEG induction in extrahippocampal structures behavioral outcomes of optogenetic manipulations of hippo-
following artificial activation of engram-bearing hippocampal campal engrams appear experience dependent. For example,
cells does not necessarily indicate that these are the precise ex- as noted above, stimulation of contextual fear-conditioning-
trahippocampal neurons involved in the original learning. More- tagged cells in the DG results in increased freezing in a safe
over, specific controls, such as untagged, context-exposed (no shock) context (e.g., Liu et al., 2012). However, if such stim-
and nonreinforced animals, are often essential to address issues ulation of the DG occurs for neurons that were tagged following
of memory versus performance; indeed, animals may be able to the extinction of fear in a shock-associated context, then this
use alternative learning or generalization strategies to achieve manipulation results in reduced freezing in a shock-associated
task-dependent behavior (e.g., Wiltgen et al., 2010), even in context and decreased spontaneous recovery of contextual
the absence of the hippocampus (for discussion, see Maren fear (Lacagnina et al., 2019). Likewise, inhibition of context-
et al., 2013). So, what is the evidence for learning-specific and fear-tagged DG cells attenuates freezing in a shock-associated
hippocampus-dependent reinstatement of neural activity? To context (Tanaka et al., 2014), but inhibition of extinction-tagged
this end, one study has shown that photoinhibition of learning- DG cells can increase defensive responding in a previously extin-
tagged CA1 pyramidal cells resulted in the reduction of fear guished context (Lacagnina et al., 2019). These experience-spe-
behavior in a shock-associated context and that this coincided cific findings are complemented by other studies where the
with reduced reinstatement of c-Fos expression, specifically in behavioral response (beyond defensive behavior) of DG engram
other c-Fos-tagged and, presumably, engram-bearing cortical reactivation reflects the valence of the reinforcing stimuli associ-
and subcortical cells of the brain (Tanaka et al., 2014). In a sepa- ated the context or engram (e.g., Ramirez et al., 2015; Redondo
rate study (Guo et al., 2018), it was found that contextual fear et al., 2014). Also, consider that simply reactivating engram cells

4 Neuron 107, September 9, 2020


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

that were tagged during homecage exploration or exposure to a patterns. Indeed, one study found that reactivation of cells of the
context in which shock never occurred does not appear to RSC that were tagged at conditioning is sufficient to induce
induce abnormal locomotion or overt defensive behaviors (e.g., behavioral expression of fear, even in the absence of a fully func-
Ghandour et al., 2019). tional hippocampus (Cowansage et al., 2014). Importantly, this
While indexing may explain the capacity for photoactivation of study showed that optogenetic activation of the RSC engram,
tagged hippocampal engram cells to trigger memory-specific like natural recall, recruited overlapping downstream circuits in
behavior in different contexts, reinstatement of behavior is often the amygdala and EC, demonstrative of reinstatement of experi-
notably less than what would be expected through natural recall ential activity. Thus, indexing may not necessarily be unique to
(i.e., returning the animal to the original training context). In the the hippocampus; however, the hippocampus may be uniquely
framework of indexing, we propose there are number of reasons positioned to index episodic events, given the significantly
why this may be the case beyond the fact that natural recall is greater extent to which it integrates complex and hierarchical
presumably most effective in reactivating the index. Importantly, sensory information from across the brain (see sections below),
the abovementioned tagging systems, while experimentally time as well as due to its discriminative coding and circuit architec-
locked, are still thought to open a window of tagging that may be ture. Unpublished findings indicating that reactivation of
on the order of at least several hours. Thus, when artificially reac- engram-tagged neural structures, outside the hippocampus,
tivating these cells, it is possible that the experimenter may also does not equally reinstate behavior may support the particular
be triggering activation of other nonspecific indices and/or importance of hippocampal indexing (Roy et al., 2019).
experiences such as activity in the homecage and pre- or Memory Indexing in Humans?
post-training handling. These patterns are not specific to the pri- Electrophysiological studies in humans have suggested that the
mary learning episode in question and thus may compete for human hippocampus also possesses properties consistent with
behavioral expression. Contextual stimuli present in the test indexing. For example, in epileptic patients with depth elec-
context may also trigger interference as well, acting as external trodes implanted into the medial temporal lobe, free recall of
inhibitors. Thereby, a number of controls (e.g., nonreinforced, an audiovisual experience was shown to follow the selective re-
homecage) for better isolating and assessing the degree of activation of hippocampal and EC cells that were active during
experienced-dependent behavioral reinstatement should be the prior experience (Gelbard-Sagiv et al., 2008). Likewise,
performed. Additionally, hippocampal indices not only are pro- successful retrieval in an object association task coincided
posed to encode the relevant brain systems activated during with reinstatement of spiking activity in hippocampal and EC
an experience but also may represent the sequential patterns cells, hippocampal activity preceding EC firing, and decoding
of such activation (Buzsáki and Tingley, 2018). Current methods analyses of EC activity predicting the identity of the recalled ob-
of optogenetic reactivation of hippocampal engram cells lack ject (Staresina et al., 2019). Other intracranial recordings have
such sequence-based reactivation (Carrillo-Reid et al., 2019), shown that behavioral recall was linked to coordinated hippo-
beyond what is inherently structured in the linkage of hippocam- campal-lateral temporal cortical representational reinstatement
pal engram-bearing circuits. Further technological advance- of item-context associations (Pacheco Estefan et al., 2019). In
ments, which may better constrain the window of tagging to a this study, hippocampal reinstatement preceded that seen in
particular experience or may be able to reactivate cells in the neocortex, and hippocampal-cortical gamma phase syn-
sequence-dependent manners, are crucial to improve the read- chrony during hippocampal reinstatement predicted neocortical
outs and interpretation of this reinstatement. reinstatement. Moreover, these findings are mirrored in addi-
Is the hippocampus alone in its potential capacity for index- tional studies that have found memory-related reinstatement in
ing? Association cortices such as the sensory association cortex the human hippocampus is underscored by a sparse and distrib-
and EC may also exhibit indexing properties due to their conver- uted set of active cells (Wixted et al., 2014, 2018).
gence of sensory input, thereby contributing to a hierarchical in- Human functional magnetic resonance imaging (fMRI) studies
dexing scheme (McClelland et al., 1995). Thus, the hippocampus support a similar interpretation. For example, one fMRI study
may serve to some extent as an index of indices in the EC and (Harand et al., 2012) reported that hippocampal BOLD (blood-
other input structures as information is routed in and then back oxygen-level-dependent) activity during an episodic learning
out again. Assuming such hippocampal signals can be decom- experience matched its activity at recall (i.e., recognition of pre-
pressed to reinstate activity in cortical and subcortical nuclei viously shown visual cues), particularly when subjects reported
(as noted above), an exact one-to-one representation in the hip- the remembering of episodic details of the learning event. Inter-
pocampus of cortical modules (for example) seems unlikely and estingly, this episodic reinstatement of hippocampal activity
may not be necessary. In fact, convergence of neural activity into occurred for remembered cues at 3 days and even 3 months
the hippocampus might be essential for its abilities to form following learning. Moreover, for successful retrieval of experien-
conjunctive contextual representations (Rudy and O’Reilly, tial memory (in tasks such as object recall and recognition),
1999). Other critical targets of the hippocampus, such as the ret- regions including the RSC, parahippocampal cortex (PHC), peri-
rosplenial cortex (RSC) (Cowansage et al., 2014; Mao et al., rhinal cortex (PRC), and prefrontal cortex (PFC) have all been
2018) or lateral septum (LS) (Bender et al., 2015; Besnard shown to exhibit recall-dependent reinstatement along with or
et al., 2019; Tingley and Buzsáki, 2018), may also maintain in close temporal proximity to hippocampal reinstatement, sug-
such convergence of processing and may thereby be part of a gestive of hippocampal-dependent routing (e.g., Arnold et al.,
hierarchical indexing scheme, assuming these structures are 2018; Jonker et al., 2018; Schultz et al., 2019). Again, while rein-
capable of reinstating patterns of experience-specific assembly statement and temporal patterns of activation alone do not

Neuron 107, September 9, 2020 5


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

An Integrated Circuit Model of the Hippocampal Index


In the decades since the introduction of the hippocampal mem-
ory indexing theory, considerable advances have been made in
our understanding of the complexity and diversity of hippocam-
pal circuits. If experience-tagged hippocampal engrams serve
as episodic indices, how might the circuit architecture of the
brain be employed for encoding and recall? To these ends,
the conjunctive, sparse, and compressed code generated in
the DG via pattern separation would support indexing by mini-
mizing memory interference (Figure 2, feature i) (Cayco-Gajic
and Silver, 2019; Hainmueller and Bartos, 2020; Knierim and
Neunuebel, 2016; McHugh et al., 2007), while DG outputs,
together with direct EC inputs (alongside the diverse afferents
described below), onto CA3 cells would bias attractor dynamics
in the recurrent network to store an experience as a new mem-
ory, or catalyze the retrieval or updating of a previously encoded
memory by pattern completion (Figure 2, features ii–iv). Accord-
ingly, the experience is registered in a sparse hippocampal code
or engram composed of principal cells (and inhibitory neurons
[INs]) across the different subregions (DG, CA3, CA2, and
CA1), with their coordinated activity permitting intra- and extra-
hippocampal reinstatement of the original experience through
dynamic routing (Figure 2, features v–vii; Figures 3 and 4). We
elaborate on this idea with recent examples in the next sections.
Dynamic Routing: Hippocampal Afferents
In this framework, the DG-CA3-CA2-CA1 circuit can be perceived
as a template of nodes, with each node receiving diverse intra- and
extrahippocampal inputs allowing for the integrative, dynamic,
and flexible incorporation of cognitive and visceral information
into memory representations (Figure 3A). These properties would
enable the hippocampus to participate in many ‘‘types’’ of mem-
ories—spatial, goal-oriented, social, future-planning—all which
may comprise diverse episodic experiences. For example, direct
long-range GABAergic projections from the lateral hypothalamus
(LH) to CA3 have been recently identified (Zhou et al., 2019a);
these neurons synapse onto CA3 interneurons and appear to
have critical functions in tasks of object recognition and discrimi-
nation, revealing a direct pathway by which CA3 may integrate
endocrine signals in learning and memory processes. CA3 neu-
rons also incorporate locus coeruleus (LC) input, and one study
found that LC projections to dorsal CA3 (dCA3; but not to CA1
or DG) are required for encoding (but not retrieval, which may be
Figure 3. Hippocampal (DG-CA3-CA2-CA1) Circuit Architecture and
Anatomical Loops Permit Flexible Integration and Routing of mediated by CA3-CA1 [see below]) of a contextual representation
Experiential Information (as assessed by distance traveled in a previously explored context
(A) Examples of hippocampal inputs (blue arrows). or via single-trial contextual fear conditioning) (Wagatsuma et al.,
(B) Examples of hippocampal outputs (orange arrows). Note that the pro-
jections shown are not exhaustive.
2018). Additionally, parallel circuits projecting from neurons in
Brain regions include the anterior cingulate cortex (ACC), bed nucleus of the the supramammillary nucleus (SUM) to the DG and CA2 have
stria terminalis (BNST), basolateral/basomedial amygdala (BLA/BMA), central been found to carry contextual and social novelty signals, respec-
amygdala (CEA), dentate gyrus (DG), dorsal raphe nucleus (DRN), entorhinal tively, allowing hypothalamic sculpting of hippocampal memory in
cortex (EC), cornu ammonis regions (CA1–CA3), infralimbic cortex (IL), locus
coeruleus (LC), anterior/lateral hypothalamic area (A/LHA), lateral septum (LS), a task-specific manner (Chen et al., 2020; see also Li et al., 2020;
medial septum (MS), nucleus accumbens (NAC), orbital frontal cortex (OFC), Hashimotodani et al., 2018). These recent discoveries broaden our
prelimbic cortex (PL), nucleus reuniens (RE), retrosplenial cortex (RSC), sub- understanding of the diversity of mammalian hippocampal affer-
iculum (SUB), supramammillary nucleus (SUM), and ventral tegmental area
(VTA). Brain region images were generated using Brain Explorer 2.0 (Lein
ents and point to multiple sources via which the hippocampus
et al., 2007). may integrate signals for memory formation or recall. Activity of
these distinct sets of inputs (alongside other important inputs,
demonstrate indexing, these findings are consistent with data including from the amygdala and anterior cingulate cortex), re-
from rodents and leave open the possibility that future experi- cruited based on ongoing experience, may govern which hippo-
ments may directly test this idea in humans. campal routes are deployed for encoding and/or recall.

6 Neuron 107, September 9, 2020


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

Dynamic Routing: Hippocampal Efferents updating of a reward location promoted activity remapping in
Further arguing against a simple sequential processing loop, it is CA1 and CA3, but not in the DG (Hainmueller and Bartos,
clear that each node within the DG-CA3-CA2-CA1 circuit pro- 2018). Recent work has also demonstrated that contextual fear
jects to distinct outputs (Figure 3B). In the framework of index extinction recruits a distinct DG engram from that encoding the
theory, these outputs may be flexibly deployed by engram cells original context-shock association, which reduces levels of
to reinstate an experience (whether that experience is appetitive, freezing in the training context but can be overcome by the orig-
aversive, etc.). An example of this potential selective routing can inal engram to induce relapse (Lacagnina et al., 2019). Likewise,
be found in ventral CA1 (vCA1) neurons (Ciocchi et al., 2015). In prolonged optogenetic or chemogenetic reactivation of a DG
this study, vCA1 cells were tracked based on their projections to fear memory engram in the training context without the uncondi-
the PFC, nucleus accumbens (NAc), and amygdala during tioned stimulus promoted extinction of the conditioned response
behavior, revealing that activity in these neurons were task and (Khalaf et al., 2018). Of course, the hippocampus is not unique in
pathway dependent. These findings critically suggest that sig- this type of broad connectedness; other hubs in the brain,
nals out of CA1 are not uniformly transmitted to its targets; including the claustrum (Jackson et al., 2020) and the thalamus
rather, it supports the idea of that efferent hippocampal signals (Halassa and Sherman, 2019), may surpass it in terms of total
are routed based on task and mnemonic demands. With partic- connectivity. However, these data suggest that the hippocam-
ular relevance to indexing, dorsal CA1 (dCA1) tetrode recordings pus is capable of integrating and routing complex information
paired with circuit-specific optogenetics have shown that from a variety of source structures, supporting its role in the bind-
expression of conditioned place preference (CPP) depends on ing of cognition and emotion to subserve memory (Figure 4A).
the reinstatement of dCA1 representations that were active dur- Dynamic Routing: Inhibitory Microcircuits
ing training (Trouche et al., 2019). Furthermore, CPP expression How might these diverse communication channels running
is lost if dCA1 terminals in NAc are photoinhibited, despite dCA1 through the hippocampus be managed? Hippocampal INs are
pyramidal cells maintaining their context-dependent cell assem- well positioned to function as arbiters of information flow in the
blies during testing (see also Zhou et al., 2019b). hippocampus, as they target different cellular compartments of
This routing ability is not restricted to CA1; CA3 output neu- principal neurons, are reciprocally connected with other interneu-
rons may route information via projections to CA1, CA2, or the rons, and project locally within and across different subregions
DLS (dorsolateral septum). Indeed, brain-wide analyses of co- and lamellae and out of the hippocampus to association cortices
activated circuits accompanying contextual fear discrimination and subcortical circuits (long-range INs) (Caroni, 2015). Moreover,
identified a CA3-DLS module (Besnard et al., 2019). This hippocampal INs modulate neuronal excitability, summation of
pathway appears to recruit somatostatin (SST)+ DLS cells to excitatory inputs, and neuronal firing in addition to generation of
gate conditioned freezing, as in vivo calcium imaging found network oscillations (theta and gamma oscillations) and as such
SST+ DLS cell activity reliably discriminated shock-associated are thought to play critical roles in local circuit computations un-
versus safe contexts. In support of these findings, optoge- derlying exploration and encoding, action selection, memory
netic-terminal-specific silencing of dCA3 terminals in dCA1 consolidation, retrieval, and reinstatement (Cardin, 2018; Makino
and DLS has suggested distinct roles for dCA3-CA1 and et al., 2019; Roux and Buzsáki, 2015; Sosa et al., 2018). Indeed,
dCA3-DLS projections to contextual fear learning (or consolida- recent studies have uncovered a diverse population of INs in
tion) and discrimination, respectively (Besnard et al., 2020). CA1 and CA3 that exert perisomatic and dendritic inhibition on
For CA2, its efferent network positions it strongly for memories DGCs and are modulated by SWRs.
involving social recognition, discrimination, and aggression (Mid- Local INs may regulate information flow within a hippocampal
dleton and McHugh, 2019). Indeed, CA2 (and CA3) efferents do subregion by biasing recruitment of principal cells, thereby
not uniformly regulate discrimination (Raam et al., 2017). Optoge- creating parallel channels as evidenced in a study that identified
netic experiments have demonstrated that anterior CA2/dCA2 biased PV+ basket cell (BC) connectivity with deep and superfi-
neurons targeting dCA1 are essential for novel object recognition, cial CA1 neurons of the ventral hippocampus (Lee et al., 2014).
but not for discrimination between novel and familiar conspecifics. The authors found that PV+ BCs preferentially innervated deep
The opposite was true for dCA2/dCA3 projections to posterior CA1 pyramidal neurons but received greater excitatory inputs
CA1. Photoinhibition of dCA2/dCA3 projections to the DLS were from superficial CA1 pyramidal neurons. At the level of output,
instead shown to somewhat enhance social discrimination, but PV+ BCs exert greater inhibition onto basolateral amygdala
with no effect on object discrimination or recognition. In social be- (BLA)-projecting deep CA1 neurons than those that projected
haviors, axons from dCA2 neurons targeting ventral hippocampus to the PFC and, in turn, received greater excitatory input from
were shown to be critically involved in maintaining memory of a PFC than BLA. These data suggest that PV+ BCs do not uni-
familiar animal (Meira et al., 2018; Raam et al., 2017), while phar- formly inhibit CA1; instead, it is likely that PV+ BC-principal cell
macogenetic inhibition of CA2 terminals in the DLS attenuates so- microcircuits bias information flow to distinct vCA1 outputs,
cial aggression (Leroy et al., 2018), a pathway that, when active, including PFC, BLA, NAc, DLS, and LH (serving dynamic and
appeared to invoke DLS-innervation of the ventromedial hypothal- flexible routing). Local INs may also differentially regulate theta
amus to drive attack behavior. phase-locking and burst firing of CA1 neurons through somatic
In total, multiple nonoverlapping engrams within these diverse or dendritic inhibition, respectively (Royer et al., 2012). Because
hippocampal routes may compete through updating or ongoing burst firing of pyramidal cells is thought to increase synaptic
learning to modify behavioral output. For example, two-photon communication by increasing excitation of downstream targets,
(2P) imaging of DG and CA3 engrams in vivo revealed that the local INs may modulate CA1 outputs through this mechanism

Neuron 107, September 9, 2020 7


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

A Figure 4. Hippocampal Engrams May Index


Experience to Reinstate Experiential
Memory
(A) Distinct experiences are thought to be encoded
within DG-CA3-CA2-CA1 connections, with
engram-bearing cells being functionally linked to
other neurons for the same episode. Recall can
then be driven by partial input (cues) that reinstate
activity (filled-in circles/squares) within hippo-
campal circuits to drive extrahippocampal rein-
statement via its diverse outputs and connectivity
to other engram-bearing cells. PCs, principal cells;
INs, inhibitory neurons.
(B) Hippocampal cells may register more than one
experience in distinct patterns of connectivity
prescribed by activity-dependent gene expres-
sion (shown here as combinations of 1 s and 0 s).

press CA1 cholecystokinin-positive


(CCK+) interneurons that relay feed-for-
ward inhibition from CA3 to CA1 ex vivo
B (Basu et al., 2016). This disinhibition of
CA1 interneurons induced enhanced den-
dritic spiking within a specific temporal
window, a mechanism by which sensory
information and mnemonic information
arriving from excitatory LEC inputs and
CA3, respectively, may be integrated.
More recently, a class of long-range inhib-
itory neuronal nitric oxide synthase
(nNOS)-expressing cells in CA1 (LINC neu-
rons) has been identified that project both
locally and extra-hippocampally (Christen-
son Wick et al., 2019). These neurons
inhibit superficial and deep principal cells
and other INs in CA1 and project
to diverse extrahippocampal targets,
including the tenia tecta, subiculum, hypo-
thalamus, olfactory bulb, and EC. Optoge-
netic activation of LINC neurons entrained
hippocampal oscillations and hippocam-
pal-frontal cortex (tenia tecta) coherence.
(Graves et al., 2012; Lisman, 1997; Takahashi and Magee, 2009). Thus, converging evidence has begun to illuminate how cell phys-
Importantly, rhythmic optogenetic activation of PV+ INs in CA1 to iology, activity-dependent gene expression, and microcircuit con-
mimic that seen during learning enhanced theta, delta, and ripple nectivity support hippocampal engram-cell-dependent indexing
oscillations; stabilized functional connectivity between CA1 neu- (i.e., encoding of experiences and routing of information to mediate
rons; and reliably promoted ensemble reactivation (Ognjanovski reinstatement). We discuss these features of engram cell iden-
et al., 2014). The exact role these oscillations play in the ability of tity next.
an index to reactivate downstream targets remains largely un-
tested; however, evidence suggests the coherence or coordina- Index Cell Identity
tion of activity they provide may facilitate both the encoding and Given the long-standing focus on rodent hippocampal place
recall of memories across various structures by ensuring tempo- cells, an obvious question is what aspect of contextual memory
ral coordination of activity (Buzsáki, 2015; Corcoran et al., 2016; is encoded within the hippocampal engram. Behavioral studies
Igarashi et al., 2014; Joo and Frank, 2018; Lin et al., 2017; Ma- using variations of contextual fear conditioning suggest that
kino et al., 2019; Wirt and Hyman, 2019). the hippocampus generates a conjunctive representation of
Pioneering in vivo recordings and imaging studies in rats identi- multimodal sensory information formed through physical explo-
fied extensively connected INs with extrahippocampal (septum, ration of a context (Fanselow, 2000; see also Krasne et al., 2015).
subiculum, para- and pre-subiculum, and RSC) projections that For example, one study preexposed rats to either the condition-
coordinate network oscillations (Bonifazi et al., 2009; Jinno, ing context or independent features of that context and found
2009). Long-range inhibitory projection neurons of the LEC sup- context fear after an immediate shock is facilitated only when

8 Neuron 107, September 9, 2020


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

these multimodal cues are presented together, suggesting that average firing rates, which were highly correlated between the first
the hippocampus represents the conjunction of the cues and second visits, while the firing rates in a distinct context were
defining the context (Rudy and O’Reilly, 1999). Indeed, tempo- strongly altered. It worth noting that this strong correlation of ac-
rary pharmacological inactivation of the hippocampus during tivity emerges as soon as animal is placed in the environment,
the context preexposure prevents the contextual fear condition- suggesting their activity could support rapid retrieval of contextual
ing of immediate shock (Matus-Amat et al., 2004). Past studies of memory, consistent with the indexing theory. Further, these find-
IEG expression in the hippocampus support this this view (e.g., ings of a unique physiology suggest the importance of the tempo-
Huff et al., 2006; Zhu et al., 1997); the strongest IEG response ral relationship between input from CA3 and the EC in triggering
is achieved when a novel combination of multimodal cues is pre- CA1 pyramidal cell plasticity and activity in vivo (Bittner et al.,
sented to the animal. Conversely, hippocampal IEG expression 2017; Ketz et al., 2013).
is weak or nonexistent when a highly habituated stimulus is Complementary results were also found in a physiological study
given. Note that immediate shock upon context entry in the in which c-Fos-positive CA1 neurons were inhibited during recall
absence of preexposure (and extensive postexposure) does (Trouche et al., 2016). In this study, engram neurons, defined by
not appear to elevate levels of IEGs in the hippocampus relative c-Fos expression associated with the acquisition of a cocaine-re-
to a habituated homecage (see also Erwin et al., 2020). These warded CPP, were labeled with an inhibitory opsin. Inactivation of
data suggest that IEG-expressing engram neurons may not this ensemble during a subsequent recall session reduced CPP
necessarily or exclusively store spatial information, as rodents behavior and, interestingly, led to a global remapping of the
will have active place cells even in the most familiar of contexts, c-Fos-negative active place cell population. Together, these
but rather hippocampal circuits detect novelty in the combina- data suggest that the role of the c-Fos-positive place cells is to
tion of sensory cues and encode it as a contextual representation serve as a context-specific memory index and that their activity
supporting the episodic experience. Indeed, optogenetic stimu- is crucial for the stable reinstatement of a more detailed spatial
lation of CA1 neurons tagged in a novel context the day prior to map, consisting of the remainder of the place cell population,
immediate shock delivery in the same context, but not a different that would permit animals precise navigation.
context, resulted in retrieval of the contextual fear memory While it may seem odd at first that the neurons indispensable
(Ghandour et al., 2019; see also Ramirez et al., 2013). Thus, a for inducing memory recall in CA1 show spatial instability, recent
memory engram, defined by active principal cells during contex- computational modeling lends support to this view (Benna and
tual learning, may preferentially encode conjunctive contextual Fusi, 2019). Based on an assumption that the hippocampus en-
information, as opposed to specific locations that could be codes correlations of incoming sensory information, similar to
biased by a specific cue or subset of cues. the interpretation from contextual conditioning and IEG studies,
Physiology the model predicted instability in the spatial representation of
Key insights into the precise identity of engram-bearing cells the hippocampus. Their ultrametric tree-like network generated
came from in vivo recording of CA1 neurons in a mouse in which sparse and compressed representations of inputs to efficiently
c-Fos-positive neurons labeled during a novel context exposure store uncorrelated patterns in a hippocampal-like network.
were tagged with channelrhodopsin 2 (ChR2) and subsequently When spatial navigation in a 2D open field is simulated, activities
optically identified (Tanaka et al., 2018). As expected, 50% of of hippocampal cells in the model exhibit strong modulation by
all CA1 pyramidal cells could be classified as place cells; however, the animal’s location within the environment (i.e., place cells).
only one-quarter of these place cells also expressed c-Fos (opto- However, similar to the experimental observations above, these
genetically identified). In short, engram cells were place cells, but place fields significantly remapped between epochs in the
only one-quarter of place cells were engram cells. During memory same environment, suggesting instability of firing location as a
encoding, these engram-bearing (c-Fos) cells are distinguished by reflection of correlation coding rather than spatial coding. Taken
higher mean firing rates (as also seen in a calcium imaging study; together, these studies support a view that activity of the hippo-
Ghandour et al., 2019), repetitive bursts of action potentials at the campal engram reflects more than just space and suggest at least
theta frequency, and higher entrainment by the local fast gamma a subset of neurons are dedicated to capturing the conjunctive
oscillation compared to the non-c-Fos-expressing place cells, correlations that define the larger context of the experience.
again highlighting the role inhibitory circuits and oscillations may Activity-Dependent Regulation of Gene Expression and
play in the formation of the index. Interestingly, when mice were Connectivity
returned to the context the next day, c-Fos-positive engram neu- Hippocampal engrams are generated and maintained by
rons, while remaining place cells (meaning they still demonstrated strengthening or modification of synapses among activated cells
a reliable in-session spatially receptive field), showed a much within and across subregions. One study found that c-Fos-tagged
higher degree of spatial instability compared to the encoding ses- CA3 cells preferentially responded to stimulation of engram-
sion (remapping) than the c-Fos-negative place cell population. tagged DGCs (Ryan et al., 2015), results indicative of experi-
These data can be seen as paradoxical; how is it that the neurons ence-driven connectivity of DG-CA3. Likewise, context-fear-
shown to be capable of reinstating context-appropriate behavior dependent increases in the number and size of spines in
show relatively lower spatial specificity than the remaining active engram-bearing cells of CA1 coincide with direct input from
cells? Importantly, when only firing rate (and not location) was engram-tagged cells from CA3 (Choi et al., 2018). Additionally,
considered, it was clear that the engram cells faithfully encoded DG engram cells were shown to exhibit greater connectivity
contextual identity, but not specific location. A return to the en- with stratum lucidum PV+ INs than non-engram DG cells (Guo
coding context resulted in c-Fos-positive neurons reinstating their et al., 2018). These observations have motivated investigations

Neuron 107, September 9, 2020 9


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

into how developmental programs and activity-dependent gene dexes acquire this function over time? Is the cortical index equiv-
expression prescribe engram formation. First, principal neurons alent to the hippocampal index?
may have differing propensities toward recruitment into engrams Several lines of evidence support the notion that the hippo-
based on developmentally programmed intrinsic firing properties campal index may be necessary for maintenance and retrieval
and connectivity (Cembrowski and Spruston, 2019; Soltesz and of only highly precise memories. First, although hippocampal
Losonczy, 2018). Second, activity-dependent transcription fac- damage at remote time points may still permit retrieval of
tors and combinations thereof enable neurons to read detailed contextual representations, the extent of memory
patterns of neural activity and transcribe molecular specifiers of retrieval is often much less robust (Wang et al., 2009), suggesting
connectivity to facilitate strengthening or modification of synap- that extrahippocampal indices may not fully compensate for the
ses (Tyssowski et al., 2018). For example, cyclic adenosine mono- loss of the hippocampal index. Indeed, while a similar degree of
phosphate (cAMP)-response-element-binding protein (CREB)- hippocampal activation is seen for recent and remote memories,
overexpression studies have revealed that enhancing basal the reactivation patterns are different (Tayler et al., 2013; see also
activity and excitability of principal cells in the hippocampus (or Guskjolen et al., 2018). Second, artificially stabilizing the engram
subsets of amygdalar neurons, etc.) prior to learning can bias within DG-CA3 decreases remote memory generalization (and
the allocation and tagging process to these cells without altering maintains behavioral reinstatement of remote DG engram stimu-
the overall size of the engram per se (Josselyn and Frankland, lation; Guo et al., 2018), providing a direct link between mainte-
2018; Josselyn and Tonegawa, 2020). A recent report using nance of the hippocampal index and remote memory precision.
RNA sequencing of engram-tagged DG cells (following contextual However, maintenance of separate hippocampal indices for all
fear conditioning) identified a unique learning-dependent genetic episodic memories is thought to require significantly greater ca-
profile for engram-bearing cells, with CREB-dependent transcrip- pacity than is available to avoid memory interference (McClel-
tion networks being differentially regulated and required for land et al., 1995; Miller and Sahay, 2019; Skaggs and McNaugh-
consolidation (Rao-Ruiz et al., 2019). Interestingly, many of these ton, 1992). In CA1, the various methods employed to genetically
genes were previously shown to regulate somatic inhibition (e.g., label engram neurons typically capture 20% of the pyramidal
neuronal PAS domain protein 4 [NPAS4], proenkephalin [Penk], cells in the region (Tanaka et al., 2018), and in vivo imaging sug-
and brain-derived neurotrophic factor [BDNF]) (Bloodgood gested that the identity of these allocated neurons shifts over the
et al., 2013). Consistent with these findings, within the DG, timescale of hours (Cai et al., 2016); thus, it appears that natural
contextual fear learning regulates CREB-dependent levels of neu- decay of hippocampal indices ensures the time-dependent reor-
ropeptide Y (NPY) in SST+ hilar perforant path-associated (HIPP) ganization of memory traces to support different degrees of
interneurons, which may regulate SST+ HIPP-mediated feedback generalization and generation of schema to facilitate new
and feed-forward inhibition in the DG to govern the size of the learning. It may be that some hippocampal indexes, perhaps
engram (Raza et al., 2017; Stefanelli et al., 2016). Not surprisingly, for salient life events, are maintained for longer periods of time,
different IEG transcription factors (including NPAS4) have been thereby permitting recall of remote memories with high fidelity.
functionally implicated in linking principal cells with distinct IN net- The integrity and composition of the cortical indices depends
works to support engram formation (Sun et al., 2020). Thus, expe- on how competition for representation of episodic memories and
riential input may drive unique IEG expression to govern functional abstraction of statistical commonalities across ensembles
allocation of engram-bearing cells to work in concert for memory dictate the balance between preservation of details versus gen-
expression (Figure 4B). eration of schema to facilitate memory generalization. This may
involve time-dependent changes in the exact number of cells
Hippocampal Index Stability and Memory Fidelity and patterns of efferent connectivity of cortical ensembles and
Memory consolidation is thought to involve transformation and linkage of distinct engrams of separate experiences via some
reorganization of hippocampal-linked cognate cortical represen- degree of overlapping and synchronous activation (during recall
tations and a gradual decay of the hippocampal engram over or reconsolidation) of engram-bearing cells (Abdou et al., 2018;
time (DeNardo et al., 2019; Guskjolen et al., 2018; Kitamura DeNardo et al., 2019; Kitamura et al., 2017; Ohkawa et al.,
et al., 2017; Roy et al., 2017; Tayler et al., 2013; Winocur et al., 2015; Oishi et al., 2019; Pignatelli et al., 2019; Ramirez et al.,
2007). Consolidated memories have been shown to generalize 2013; Redondo et al., 2014).
or lack detail, including the extent to which they may elicit While no one model can explain all the current data, from the
visceral or physiological reactions, leading to the suggestion perspective of engrams and indexing, we favor the hypothesis
that the role the hippocampus plays in memory is to contribute of a time-dependent shift in the indexing function from the hippo-
episodic detail (Yonelinas et al., 2019). If this contribution relies campus to cortical traces concurrent with a silencing or loss of
on the initial memory trace or not remains a contested topic. the original hippocampal index (Tonegawa et al., 2018). Ulti-
For example, it has been argued that hippocampal memory mately, time-dependent shifts in hippocampus-dependent
traces remain, even for older memories (Moscovitch and Nadel, episodic detail may be useful in the development of experiential
2019), but others have argued there is a shift in the role the hip- schemas and broader knowledge.
pocampus from one of recall to reconstruction in the absence of
the original trace, with the activity serving to index consolidated Hippocampal Dysfunction
neocortical traces (Barry and Maguire, 2019a, 2019b). These Indexopathies
observations raise the following questions: Is the hippocampal Memory deficits and hippocampal dysfunction accompany
index always necessary for memory retrieval, or do cortical in- traumatic brain injury, epilepsy, age-related cognitive decline,

10 Neuron 107, September 9, 2020


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

Figure 5. Outstanding Challenges for Hippocampal Indexing and Memory Research

Alzheimer’s disease (AD), and numerous other psychiatric disor- such circuit loss, AD is considered, in part, a disease of memory
ders, including posttraumatic stress disorder (PTSD) and schizo- retrieval failure (Leal and Yassa, 2018; Roy et al., 2016).
phrenia (Besnard and Sahay, 2016; Haberman et al., 2017; Small Promoting Indexing: New Directions in Therapy
et al., 2011). Can these disorders of experiential memory be clas- Recent advances in technology and medicine have led to a num-
sified as ‘‘indexopathies,’’ insofar as they are marked by an ber of new therapeutic avenues for memory disorders—strate-
inability to accurately or precisely encode or effectively gies that may be effective, in part, because they promote or
implement hippocampus-dependent routing of information reestablish hippocampal indexing functions. For example,
(i.e., indexing)? For example, recent work documented declining growing interest has centered on deep-brain and closed-loop
hippocampal and cortical reinstatement in aging individuals feedback neuroprosthetics for symptom management in individ-
(Trelle et al., 2020). Disease- or aging-induced excitation-inhibi- uals when other lines of treatment have failed (Grosenick et al.,
tory imbalance in hippocampal circuits, which may impede in- 2015; Takeuchi and Berényi, 2020). Perhaps by restoring
dexing, may underlie much of these memory dysfunctions. context-dependent routing (and thereby indexing), these real-
Indeed, excitation-inhibition imbalance (hypo- or hyperactivity) time electrophysiological (or opto- or chemogenetic, potentially)
at the level of CA1 (e.g., Oh et al., 2013) and CA3 (e.g., Simkin methods could dynamically normalize aberrant activity or restore
et al., 2015; Wilson et al., 2005) and loss of feed-forward inhibi- cell excitability in perturbed brain circuits (e.g., in hippocampal-
tion in DG-CA3 (e.g., Guo et al., 2018) are associated with mem- amygdalar or hippocampal-prefrontal loops). Additionally,
ory imprecision in preclinical models of aging and memory disor- recent developments in targeted gene-editing approaches
ders. Human imaging studies have further reported similar (Knott and Doudna, 2018) may permit molecular reallocation or
activity changes (e.g., hyperactivity) of hippocampal structures respecification of connectivity aimed at promoting memory pre-
(Haberman et al., 2017), such as in presymptomatic familial AD cision and accuracy. Indeed, quieting disease-related hyperex-
(FAD) individuals (Quiroz et al., 2010) or patients with amnestic cited CA3 pyramidal cells may involve targeting feed-forward
mild cognitive impairment (Bakker et al., 2012). inhibitory mechanisms of DG-CA3 (Guo et al., 2018; Viana da
Such cellular, circuit, and network-level alterations may Silva et al., 2019) or CA2-CA3 connections (Boehringer et al.,
disrupt the balance between pattern separation and pattern 2017) or inhibiting aberrant LH-CA3 activity (Zhou et al.,
completion, thereby promoting aberrant index-dependent rein- 2019a), for example. Other pharmaco- or gene therapies that
statement and recall. What previously may have been subthresh- promote neurogenesis in aging or disease states may also exist
old to trigger CA3-dependent memory recall prior to disease may as beneficial therapeutic avenues for memory impairments
be sufficient after disease onset, thereby promoting excessive (Miller and Sahay, 2019).
reinstatement and memory expression in contexts that may Novel treatments for memory impairments may not be limited
not be optimal. For example, aberrant or excessive retrieval of to invasive techniques and may involve supplementing existing
past experiences has been suggested to underlie psychosis in procedures to best tap into the indexing properties of the hippo-
schizophrenia (Tamminga et al., 2010). Additionally, degradation campus. For example, although the use of mnemonic devices for
of flexible routing to extrahippocampal targets, due to connectiv- memory treatments is not new, recent developments in technol-
ity losses in disease or injury, may also impede the abilities of the ogy, such as augmented reality or 3D interactive environments,
hippocampus to effectively integrate cortical and subcortical in- may provide novel avenues for improving memory recall within
formation for proper encoding and retrieval. Perhaps related to and beyond the clinic. Indeed, the growing ubiquity of personal

Neuron 107, September 9, 2020 11


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective

handheld devices may make mobile reminders or mnemonic Aronov, D., Nevers, R., and Tank, D.W. (2017). Mapping of a non-spatial
dimension by the hippocampal-entorhinal circuit. Nature 543, 719–722.
cues (to facilitate reinstatement of memory) for treatments or
symptom management for memory impairments more acces- Bakker, A., Krauss, G.L., Albert, M.S., Speck, C.L., Jones, L.R., Stark, C.E.,
sible or specialized. When combined with psychological treat- Yassa, M.A., Bassett, S.S., Shelton, A.L., and Gallagher, M. (2012). Reduction
of hippocampal hyperactivity improves cognition in amnestic mild cognitive
ments in the clinic, these and the abovementioned possibilities impairment. Neuron 74, 467–474.
may yield new successes in treatment-resistant memory disor-
Barry, D.N., and Maguire, E.A. (2019a). Consolidating the case for transient
ders. Moving forward, indexing may be a useful framework for hippocampal memory traces. Trends Cogn. Sci. 23, 635–636.
improving clinical therapies for memory dysfunction.
Barry, D.N., and Maguire, E.A. (2019b). Remote memory and the hippocam-
pus: a constructive critique. Trends Cogn. Sci. 23, 128–142.
Conclusions: Moving Memory Forward
Basu, J., Zaremba, J.D., Cheung, S.K., Hitti, F.L., Zemelman, B.V., Losonczy,
Perhaps the brain’s most powerful search engine, the hippo- A., and Siegelbaum, S.A. (2016). Gating of hippocampal activity, plasticity, and
campus, sits at the center of the acquisition and recall of memory by entorhinal cortex long-range inhibition. Science 351, aaa5694.
episodic memory. While the mechanisms of how this is Bender, F., Gorbati, M., Cadavieco, M.C., Denisova, N., Gao, X., Holman, C.,
achieved have been the focus of decades of research across Korotkova, T., and Ponomarenko, A. (2015). Theta oscillations regulate the
many species and disciplines, it is often challenging to relate speed of locomotion via a hippocampus to lateral septum pathway. Nat. Com-
mun. 6, 8521.
disparate lines of inquiry. Here, we have highlighted human
and animal work based on recent genetic, physiological, Benna, M.K., and Fusi, S. (2019). Are place cells just memory cells? Memory
compression leads to spatial tuning and history dependence. bioRxiv.
anatomical, and computational approaches that together sup- https://doi.org/10.1101/624239.
port an expanded view of the hippocampal memory indexing
Besnard, A., and Sahay, A. (2016). Adult hippocampal neurogenesis, fear
theory. In particular, we argue that (1) the functional roles of pu- generalization, and stress. Neuropsychopharmacology 41, 24–44.
tative hippocampal engram cells include indexing, which may
facilitate detailed recall of episodic experiences; (2) this Besnard, A., Gao, Y., TaeWoo Kim, M., Twarkowski, H., Reed, A.K., Langberg,
T., Feng, W., Xu, X., Saur, D., Zweifel, L.S., et al. (2019). Dorsolateral septum
episodic recall is facilitated by the reinstatement of engram somatostatin interneurons gate mobility to calibrate context-specific behav-
cell activity and in their experience-sculpted connectivity; (3) in- ioral fear responses. Nat. Neurosci. 22, 436–446.
dexing may not be unique to the hippocampus, but the hippo- Besnard, A., Miller, S.M., and Sahay, A. (2020). Distinct dorsal and ventral hip-
campus may be uniquely positioned to index experiential mem- pocampal CA3 outputs govern contextual fear discrimination. Cell Rep. 30,
ory; and (4) disease of the hippocampus may impede truly 2360–2373.e5.
episodic memory by disrupting its capacity for precise Bittner, K.C., Milstein, A.D., Grienberger, C., Romani, S., and Magee, J.C.
context-specific reinstatement. Nonetheless, there remain a (2017). Behavioral time scale synaptic plasticity underlies CA1 place fields.
Science 357, 1033–1036.
number of outstanding questions for the field and for future
work (Figure 5). Future work that integrates these levels of an- Bloodgood, B.L., Sharma, N., Browne, H.A., Trepman, A.Z., and Greenberg,
M.E. (2013). The activity-dependent transcription factor NPAS4 regulates
alyses will be required to understand how the dynamics of in- domain-specific inhibition. Nature 503, 121–125.
formation flow in the hippocampal circuit contributes to the en-
coding and recall processes it supports. Boehringer, R., Polygalov, D., Huang, A.J.Y., Middleton, S.J., Robert, V., Wint-
zer, M.E., Piskorowski, R.A., Chevaleyre, V., and McHugh, T.J. (2017). Chronic
loss of CA2 transmission leads to hippocampal hyperexcitability. Neuron 94,
ACKNOWLEDGMENTS 642–655.e9.

Bonifazi, P., Goldin, M., Picardo, M.A., Jorquera, I., Cattani, A., Bianconi, G.,
T.D.G. acknowledges support from Harvard Brain Science Initiative. K.Z.T. ac- Represa, A., Ben-Ari, Y., and Cossart, R. (2009). GABAergic hub neurons
knowledges support from MEXT Grant-in-Aid for Young Scientists orchestrate synchrony in developing hippocampal networks. Science 326,
(19K16305), Grant-in-Aid for JSPS fellows (19J00974), and a Nakajima Foun- 1419–1424.
dation research grant. A.S. acknowledges support from NIH Biobehavioral
Research Awards for Innovative New Scientists (BRAINS) 1-R01MH104175, Burgess, N., and O’Keefe, J. (2011). Models of place and grid cell firing and
NIH-NIA 1R01AG048908-01A1, NIH 1R01MH111729-01, the James and Au- theta rhythmicity. Curr. Opin. Neurobiol. 21, 734–744.
drey Foster MGH Research Scholar Award, the Ellison Medical Foundation
Buzsáki, G. (2015). Hippocampal sharp wave-ripple: a cognitive biomarker for
New Scholar in Aging, the Whitehall Foundation, the Inscopix Decode Award, episodic memory and planning. Hippocampus 25, 1073–1188.
the NARSAD Independent Investigator Award, Ellison Family Philanthropic
support, the Blue Guitar Fund, a Harvard Neurodiscovery Center/MADRC Buzsáki, G., and Tingley, D. (2018). Space and time: the hippocampus as a
Center pilot grant award, an Alzheimer’s Association research grant, the Har- sequence generator. Trends Cogn. Sci. 22, 853–869.
vard Stem Cell Institute (HSCI) development grant, and an HSCI seed grant.
T.J.M. acknowledges support from MEXT Grant-in-Aid for Scientific Research Cai, D.J., Aharoni, D., Shuman, T., Shobe, J., Biane, J., Song, W., Wei, B.,
(19H05646), MEXT Grant-in-Aid for Scientific Research on Innovative Areas Veshkini, M., La-Vu, M., Lou, J., et al. (2016). A shared neural ensemble links
(17H05591, 17H05986, and 19H05233), and the RIKEN Center for Brain distinct contextual memories encoded close in time. Nature 534, 115–118.
Science.
Cardin, J.A. (2018). Inhibitory interneurons regulate temporal precision and
correlations in cortical circuits. Trends Neurosci. 41, 689–700.
REFERENCES
Caroni, P. (2015). Inhibitory microcircuit modules in hippocampal learning.
Curr. Opin. Neurobiol. 35, 66–73.
Abdou, K., Shehata, M., Choko, K., Nishizono, H., Matsuo, M., Muramatsu, S.-
I., and Inokuchi, K. (2018). Synapse-specific representation of the identity of Carrillo-Reid, L., Han, S., Yang, W., Akrouh, A., and Yuste, R. (2019). Control-
overlapping memory engrams. Science 360, 1227–1231. ling visually guided behavior by holographic recalling of cortical ensembles.
Cell 178, 447–457.e5.
Arnold, A.E.G.F., Ekstrom, A.D., and Iaria, G. (2018). Dynamic neural network
reconfiguration during the generation and reinstatement of mnemonic repre- Cayco-Gajic, N.A., and Silver, R.A. (2019). Re-evaluating circuit mechanisms
sentations. Front. Hum. Neurosci. 12, 292. underlying pattern separation. Neuron 101, 584–602.

12 Neuron 107, September 9, 2020


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective
Cembrowski, M.S., and Spruston, N. (2019). Heterogeneity within classical cell Guenthner, C.J., Miyamichi, K., Yang, H.H., Heller, H.C., and Luo, L. (2013).
types is the rule: lessons from hippocampal pyramidal neurons. Nat. Rev. Neu- Permanent genetic access to transiently active neurons via TRAP: targeted
rosci. 20, 193–204. recombination in active populations. Neuron 78, 773–784.

Chen, S., He, L., Huang, A.J.Y., Boehringer, R., Robert, V., Wintzer, M.E., Poly- Guo, N., Soden, M.E., Herber, C., Kim, M.T., Besnard, A., Lin, P., Ma, X.,
galov, D., Weitemier, A.Z., Tao, Y., Gu, M., et al. (2020). A hypothalamic novelty Cepko, C.L., Zweifel, L.S., and Sahay, A. (2018). Dentate granule cell recruit-
signal modulates hippocampal memory. Nature, in press. ment of feedforward inhibition governs engram maintenance and remote
memory generalization. Nat. Med. 24, 438–449.
Choi, J.-H., Sim, S.-E., Kim, J.-I., Choi, D.I., Oh, J., Ye, S., Lee, J., Kim, T., Ko,
H.-G., Lim, C.-S., and Kaang, B.K. (2018). Interregional synaptic maps among Guskjolen, A., Kenney, J.W., de la Parra, J., Yeung, B.A., Josselyn, S.A., and
engram cells underlie memory formation. Science 360, 430–435. Frankland, P.W. (2018). Recovery of ‘‘lost’’ infant memories in mice. Curr.
Biol. 28, 2283–2290.e3.
Christenson Wick, Z., Tetzlaff, M.R., and Krook-Magnuson, E. (2019). Novel
long-range inhibitory nNOS-expressing hippocampal cells. eLife 8, e46816. Haberman, R.P., Branch, A., and Gallagher, M. (2017). Targeting neural hyper-
activity as a treatment to stem progression of late-onset Alzheimer’s disease.
Ciocchi, S., Passecker, J., Malagon-Vina, H., Mikus, N., and Klausberger, T. Neurotherapeutics 14, 662–676.
(2015). Brain computation. Selective information routing by ventral hippocam-
pal CA1 projection neurons. Science 348, 560–563. Hainmueller, T., and Bartos, M. (2018). Parallel emergence of stable and dy-
namic memory engrams in the hippocampus. Nature 558, 292–296.
Cohen, N.J., and Eichenbaum, H. (1993). Memory, Amnesia, and The Hippo-
campal System (MIT Press). Hainmueller, T., and Bartos, M. (2020). Dentate gyrus circuits for encoding,
retrieval and discrimination of episodic memories. Nat. Rev. Neurosci. 21,
Colgin, L.L., Leutgeb, S., Jezek, K., Leutgeb, J.K., Moser, E.I., McNaughton, 153–168.
B.L., and Moser, M.-B. (2010). Attractor-map versus autoassociation based
attractor dynamics in the hippocampal network. J. Neurophysiol. 104, 35–50. Halassa, M.M., and Sherman, S.M. (2019). Thalamocortical circuit motifs: a
general framework. Neuron 103, 762–770.
Corcoran, K.A., Frick, B.J., Radulovic, J., and Kay, L.M. (2016). Analysis of
coherent activity between retrosplenial cortex, hippocampus, thalamus, and Harand, C., Bertran, F., La Joie, R., Landeau, B., Mézenge, F., Desgranges, B.,
anterior cingulate cortex during retrieval of recent and remote context fear Peigneux, P., Eustache, F., and Rauchs, G. (2012). The hippocampus remains
memory. Neurobiol. Learn. Mem. 127, 93–101. activated over the long term for the retrieval of truly episodic memories. PLoS
ONE 7, e43495.
Cowansage, K.K., Shuman, T., Dillingham, B.C., Chang, A., Golshani, P., and
Mayford, M. (2014). Direct reactivation of a coherent neocortical memory of Hashimotodani, Y., Karube, F., Yanagawa, Y., Fujiyama, F., and Kano, M.
context. Neuron 84, 432–441. (2018). Supramammillary nucleus afferents to the dentate gyrus co-release
glutamate and GABA and potentiate granule cell output. Cell Rep. 25, 2704–
DeNardo, L.A., Liu, C.D., Allen, W.E., Adams, E.L., Friedmann, D., Fu, L., 2715.e4.
Guenthner, C.J., Tessier-Lavigne, M., and Luo, L. (2019). Temporal evolution
of cortical ensembles promoting remote memory retrieval. Nat. Neurosci. 22,
Howard, M.W., and Eichenbaum, H. (2015). Time and space in the hippocam-
460–469.
pus. Brain Res. 1621, 345–354.
Denny, C.A., Kheirbek, M.A., Alba, E.L., Tanaka, K.F., Brachman, R.A., Laugh-
Huff, N.C., Frank, M., Wright-Hardesty, K., Sprunger, D., Matus-Amat, P., Hig-
man, K.B., Tomm, N.K., Turi, G.F., Losonczy, A., and Hen, R. (2014). Hippo-
gins, E., and Rudy, J.W. (2006). Amygdala regulation of immediate-early gene
campal memory traces are differentially modulated by experience, time, and
expression in the hippocampus induced by contextual fear conditioning.
adult neurogenesis. Neuron 83, 189–201.
J. Neurosci. 26, 1616–1623.
Eichenbaum, H., Otto, T., and Cohen, N.J. (1994). Two functional components
Igarashi, K.M., Lu, L., Colgin, L.L., Moser, M.-B., and Moser, E.I. (2014). Coor-
of the hippocampal memory system. Behav. Brain Sci. 17, 449–472.
dination of entorhinal-hippocampal ensemble activity during associative
Erwin, S.R., Sun, W., Copeland, M., Lindo, S., Spruston, N., and Cembrowski, learning. Nature 510, 143–147.
M.S. (2020). A sparse, spatially biased subtype of mature granule cell domi-
nates recruitment in hippocampal-associated behaviors. Cell Rep. 31, Jackson, J., Smith, J.B., and Lee, A.K. (2020). The anatomy and physiology of
107551. claustrum-cortex interactions. Annu. Rev. Neurosci. 43, 231–247.

Fanselow, M.S. (2000). Contextual fear, gestalt memories, and the hippocam- Jinno, S. (2009). Structural organization of long-range GABAergic projection
pus. Behav. Brain Res. 110, 73–81. system of the hippocampus. Front. Neuroanat. 3, 13.

Fanselow, M.S., and Dong, H.-W. (2010). Are the dorsal and ventral hippocam- Jonker, T.R., Dimsdale-Zucker, H., Ritchey, M., Clarke, A., and Ranganath, C.
pus functionally distinct structures? Neuron 65, 7–19. (2018). Neural reactivation in parietal cortex enhances memory for episodically
linked information. Proc. Natl. Acad. Sci. USA 115, 11084–11089.
Foster, D.J. (2017). Replay comes of age. Annu. Rev. Neurosci. 40, 581–602.
Joo, H.R., and Frank, L.M. (2018). The hippocampal sharp wave-ripple in
Gelbard-Sagiv, H., Mukamel, R., Harel, M., Malach, R., and Fried, I. (2008). memory retrieval for immediate use and consolidation. Nat. Rev. Neurosci.
Internally generated reactivation of single neurons in human hippocampus 19, 744–757.
during free recall. Science 322, 96–101.
Josselyn, S.A., and Frankland, P.W. (2018). Memory allocation: mechanisms
Ghandour, K., Ohkawa, N., Fung, C.C.A., Asai, H., Saitoh, Y., Takekawa, T., and function. Annu. Rev. Neurosci. 41, 389–413.
Okubo-Suzuki, R., Soya, S., Nishizono, H., Matsuo, M., et al. (2019). Orches-
trated ensemble activities constitute a hippocampal memory engram. Nat. Josselyn, S.A., and Tonegawa, S. (2020). Memory engrams: recalling the past
Commun. 10, 2637. and imagining the future. Science 367, eaaw4325.

Gonzalez, W.G., Zhang, H., Harutyunyan, A., and Lois, C. (2019). Persistence Kesner, R.P., and Rolls, E.T. (2015). A computational theory of hippocampal
of neuronal representations through time and damage in the hippocampus. function, and tests of the theory: new developments. Neurosci. Biobehav.
Science 365, 821–825. Rev. 48, 92–147.

Graves, A.R., Moore, S.J., Bloss, E.B., Mensh, B.D., Kath, W.L., and Spruston, Ketz, N., Morkonda, S.G., and O’Reilly, R.C. (2013). Theta coordinated error-
N. (2012). Hippocampal pyramidal neurons comprise two distinct cell types driven learning in the hippocampus. PLoS Comput. Biol. 9, e1003067.
that are countermodulated by metabotropic receptors. Neuron 76, 776–789.
€ff, J. (2018).
Khalaf, O., Resch, S., Dixsaut, L., Gorden, V., Glauser, L., and Gra
Grosenick, L., Marshel, J.H., and Deisseroth, K. (2015). Closed-loop and activ- Reactivation of recall-induced neurons contributes to remote fear memory
ity-guided optogenetic control. Neuron 86, 106–139. attenuation. Science 360, 1239–1242.

Neuron 107, September 9, 2020 13


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective
Kitamura, T., Ogawa, S.K., Roy, D.S., Okuyama, T., Morrissey, M.D., Smith, from the successes and failures of connectionist models of learning and mem-
L.M., Redondo, R.L., and Tonegawa, S. (2017). Engrams and circuits crucial ory. Psychol. Rev. 102, 419–457.
for systems consolidation of a memory. Science 356, 73–78.
McHugh, T.J., Jones, M.W., Quinn, J.J., Balthasar, N., Coppari, R., Elmquist,
Knierim, J.J., and Neunuebel, J.P. (2016). Tracking the flow of hippocampal J.K., Lowell, B.B., Fanselow, M.S., Wilson, M.A., and Tonegawa, S. (2007).
computation: pattern separation, pattern completion, and attractor dynamics. Dentate gyrus NMDA receptors mediate rapid pattern separation in the hippo-
Neurobiol. Learn. Mem. 129, 38–49. campal network. Science 317, 94–99.

Knott, G.J., and Doudna, J.A. (2018). CRISPR-Cas guides the future of genetic Meira, T., Leroy, F., Buss, E.W., Oliva, A., Park, J., and Siegelbaum, S.A.
engineering. Science 361, 866–869. (2018). A hippocampal circuit linking dorsal CA2 to ventral CA1 critical for so-
cial memory dynamics. Nat. Commun. 9, 4163.
Krasne, F.B., Cushman, J.D., and Fanselow, M.S. (2015). A Bayesian context
fear learning algorithm/automaton. Front. Behav. Neurosci. 9, 112. Middleton, S.J., and McHugh, T.J. (2019). CA2: a highly connected intrahippo-
campal relay. Annu. Rev. Neurosci. 43, 55–72.
Lacagnina, A.F., Brockway, E.T., Crovetti, C.R., Shue, F., McCarty, M.J., Sat-
tler, K.P., Lim, S.C., Santos, S.L., Denny, C.A., and Drew, M.R. (2019). Distinct Miller, S.M., and Sahay, A. (2019). Functions of adult-born neurons in hippo-
hippocampal engrams control extinction and relapse of fear memory. Nat. campal memory interference and indexing. Nat. Neurosci. 22, 1565–1575.
Neurosci. 22, 753–761.
Moscovitch, M., and Nadel, L. (2019). Sculpting remote memory: enduring hip-
Leal, S.L., and Yassa, M.A. (2018). Integrating new findings and examining pocampal traces and vmPFC reconstructive processes. Trends Cogn. Sci. 23,
clinical applications of pattern separation. Nat. Neurosci. 21, 163–173. 634–635.
Lee, S.-H., Marchionni, I., Bezaire, M., Varga, C., Danielson, N., Lovett-Barron, Moser, E.I., Moser, M.-B., and McNaughton, B.L. (2017). Spatial representa-
M., Losonczy, A., and Soltesz, I. (2014). Parvalbumin-positive basket cells tion in the hippocampal formation: a history. Nat. Neurosci. 20, 1448–1464.
differentiate among hippocampal pyramidal cells. Neuron 82, 1129–1144.
Nadel, L., and Moscovitch, M. (1997). Memory consolidation, retrograde
Lein, E.S., Hawrylycz, M.J., Ao, N., Ayres, M., Bensinger, A., Bernard, A., Boe, amnesia and the hippocampal complex. Curr. Opin. Neurobiol. 7, 217–227.
A.F., Boguski, M.S., Brockway, K.S., Byrnes, E.J., et al. (2007). Genome-wide
atlas of gene expression in the adult mouse brain. Nature 445, 168–176. Nadel, L., Hoscheidt, S., and Ryan, L.R. (2013). Spatial cognition and the hip-
pocampus: the anterior-posterior axis. J. Cogn. Neurosci. 25, 22–28.
Leroy, F., Park, J., Asok, A., Brann, D.H., Meira, T., Boyle, L.M., Buss, E.W.,
Kandel, E.R., and Siegelbaum, S.A. (2018). A circuit from hippocampal CA2 Nakazawa, K., Quirk, M.C., Chitwood, R.A., Watanabe, M., Yeckel, M.F., Sun,
to lateral septum disinhibits social aggression. Nature 564, 213–218. L.D., Kato, A., Carr, C.A., Johnston, D., Wilson, M.A., and Tonegawa, S. (2002).
Requirement for hippocampal CA3 NMDA receptors in associative memory
Li, Y., Bao, H., Luo, Y., Yoan, C., Sullivan, H.A., Quintanilla, L., Wickersham, I.,
recall. Science 297, 211–218.
Lazarus, M., Shin, Y.I., and Song, J. (2020). Supramammillary nucleus syn-
chronizes with dentate gyrus to regulate spatial memory retrieval through O’Keefe, J., and Dostrovsky, J. (1971). The hippocampus as a spatial map.
glutamate release. eLife 9, 9. Preliminary evidence from unit activity in the freely-moving rat. Brain Res.
34, 171–175.
Lin, J.-J., Rugg, M.D., Das, S., Stein, J., Rizzuto, D.S., Kahana, M.J., and Lega,
B.C. (2017). Theta band power increases in the posterior hippocampus predict O’Keefe, J., and Nadel, L. (1978). The Hippocampus as a Cognitive Map (Ox-
successful episodic memory encoding in humans. Hippocampus 27, ford University Press).
1040–1053.
Ognjanovski, N., Maruyama, D., Lashner, N., Zochowski, M., and Aton, S.J.
Lisman, J.E. (1997). Bursts as a unit of neural information: making unreliable
(2014). CA1 hippocampal network activity changes during sleep-dependent
synapses reliable. Trends Neurosci. 20, 38–43.
memory consolidation. Front. Syst. Neurosci. 8, 61.
Lisman, J.E., and Otmakhova, N.A. (2001). Storage, recall, and novelty detec-
Oh, M.M., Oliveira, F.A., Waters, J., and Disterhoft, J.F. (2013). Altered calcium
tion of sequences by the hippocampus: elaborating on the SOCRATIC model
metabolism in aging CA1 hippocampal pyramidal neurons. J. Neurosci. 33,
to account for normal and aberrant effects of dopamine. Hippocampus 11,
7905–7911.
551–568.

Lisman, J., Buzsáki, G., Eichenbaum, H., Nadel, L., Ranganath, C., and Re- Ohkawa, N., Saitoh, Y., Suzuki, A., Tsujimura, S., Murayama, E., Kosugi, S.,
dish, A.D. (2017). Viewpoints: how the hippocampus contributes to memory, Nishizono, H., Matsuo, M., Takahashi, Y., Nagase, M., et al. (2015). Artificial as-
navigation and cognition. Nat. Neurosci. 20, 1434–1447. sociation of pre-stored information to generate a qualitatively new memory.
Cell Rep. 11, 261–269.
Liu, X., Ramirez, S., Pang, P.T., Puryear, C.B., Govindarajan, A., Deisseroth,
K., and Tonegawa, S. (2012). Optogenetic stimulation of a hippocampal Oishi, N., Nomoto, M., Ohkawa, N., Saitoh, Y., Sano, Y., Tsujimura, S., Nishi-
engram activates fear memory recall. Nature 484, 381–385. zono, H., Matsuo, M., Muramatsu, S.-I., and Inokuchi, K. (2019). Artificial asso-
ciation of memory events by optogenetic stimulation of hippocampal CA3 cell
MacDonald, C.J., Lepage, K.Q., Eden, U.T., and Eichenbaum, H. (2011). Hip- ensembles. Mol. Brain 12, 2.
pocampal ‘‘time cells’’ bridge the gap in memory for discontiguous events.
Neuron 71, 737–749. Pacheco Estefan, D., Sánchez-Fibla, M., Duff, A., Principe, A., Rocamora, R.,
Zhang, H., Axmacher, N., and Verschure, P.F.M.J. (2019). Coordinated repre-
Makino, Y., Polygalov, D., Bolaños, F., Benucci, A., and McHugh, T.J. (2019). sentational reinstatement in the human hippocampus and lateral temporal cor-
physiological signature of memory age in the prefrontal-hippocampal circuit. tex during episodic memory retrieval. Nat. Commun. 10, 2255.
Cell Rep. 29, 3835–3846.e5.
Pastalkova, E., Itskov, V., Amarasingham, A., and Buzsáki, G. (2008). Internally
Mao, D., Neumann, A.R., Sun, J., Bonin, V., Mohajerani, M.H., and McNaugh- generated cell assembly sequences in the rat hippocampus. Science 321,
ton, B.L. (2018). Hippocampus-dependent emergence of spatial sequence 1322–1327.
coding in retrosplenial cortex. Proc. Natl. Acad. Sci. USA 115, 8015–8018.
Pignatelli, M., Ryan, T.J., Roy, D.S., Lovett, C., Smith, L.M., Muralidhar, S., and
Maren, S., Phan, K.L., and Liberzon, I. (2013). The contextual brain: implica- Tonegawa, S. (2019). Engram cell excitability state determines the efficacy of
tions for fear conditioning, extinction and psychopathology. Nat. Rev. Neuro- memory retrieval. Neuron 101, 274–284.e5.
sci. 14, 417–428.
Quiroz, Y.T., Budson, A.E., Celone, K., Ruiz, A., Newmark, R., Castrillón, G.,
Matus-Amat, P., Higgins, E.A., Barrientos, R.M., and Rudy, J.W. (2004). The Lopera, F., and Stern, C.E. (2010). Hippocampal hyperactivation in presymp-
role of the dorsal hippocampus in the acquisition and retrieval of context mem- tomatic familial Alzheimer’s disease. Ann. Neurol. 68, 865–875.
ory representations. J. Neurosci. 24, 2431–2439.
Raam, T., McAvoy, K.M., Besnard, A., Veenema, A.H., and Sahay, A. (2017).
McClelland, J.L., McNaughton, B.L., and O’Reilly, R.C. (1995). Why there are Hippocampal oxytocin receptors are necessary for discrimination of social
complementary learning systems in the hippocampus and neocortex: insights stimuli. Nat. Commun. 8, 2001.

14 Neuron 107, September 9, 2020


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective
Ramirez, S., Liu, X., Lin, P.-A., Suh, J., Pignatelli, M., Redondo, R.L., Ryan, Staresina, B.P., Reber, T.P., Niediek, J., Boström, J., Elger, C.E., and Mor-
T.J., and Tonegawa, S. (2013). Creating a false memory in the hippocampus. mann, F. (2019). Recollection in the human hippocampal-entorhinal cell cir-
Science 341, 387–391. cuitry. Nat. Commun. 10, 1503.

Ramirez, S., Liu, X., MacDonald, C.J., Moffa, A., Zhou, J., Redondo, R.L., and €scher, C., Muller, D., and Mendez, P. (2016). Hip-
Stefanelli, T., Bertollini, C., Lu
Tonegawa, S. (2015). Activating positive memory engrams suppresses pocampal somatostatin interneurons control the size of neuronal memory en-
depression-like behaviour. Nature 522, 335–339. sembles. Neuron 89, 1074–1085.

Rao-Ruiz, P., Couey, J.J., Marcelo, I.M., Bouwkamp, C.G., Slump, D.E., Ma- Strange, B.A., Witter, M.P., Lein, E.S., and Moser, E.I. (2014). Functional orga-
tos, M.R., van der Loo, R.J., Martins, G.J., van den Hout, M., van IJcken, nization of the hippocampal longitudinal axis. Nat. Rev. Neurosci. 15, 655–669.
W.F., et al. (2019). Engram-specific transcriptome profiling of contextual mem-
ory consolidation. Nat. Commun. 10, 2232. Sun, X., Bernstein, M.J., Meng, M., Rao, S., Sørensen, A.T., Yao, L., Zhang, X.,
Anikeeva, P.O., and Lin, Y. (2020). Functionally distinct neuronal ensembles
€ller, B., Demiray, Y.E., Ludewig, S.,
Raza, S.A., Albrecht, A., Çalışkan, G., Mu within the memory engram. Cell 181, 410–423.e17.
Meis, S., Faber, N., Hartig, R., Schraven, B., et al. (2017). HIPP neurons in
the dentate gyrus mediate the cholinergic modulation of background context Takahashi, H., and Magee, J.C. (2009). Pathway interactions and synaptic
memory salience. Nat. Commun. 8, 189. plasticity in the dendritic tuft regions of CA1 pyramidal neurons. Neuron 62,
102–111.
Redondo, R.L., Kim, J., Arons, A.L., Ramirez, S., Liu, X., and Tonegawa, S.
Takeuchi, Y., and Berényi, A. (2020). Oscillotherapeutics: time-targeted inter-
(2014). Bidirectional switch of the valence associated with a hippocampal
ventions in epilepsy and beyond. Neurosci. Res. 152, 87–107.
contextual memory engram. Nature 513, 426–430.
Tamminga, C.A., Stan, A.D., and Wagner, A.D. (2010). The hippocampal for-
Roux, L., and Buzsáki, G. (2015). Tasks for inhibitory interneurons in intact mation in schizophrenia. Am. J. Psychiatry 167, 1178–1193.
brain circuits. Neuropharmacology 88, 10–23.
Tanaka, K.Z. (2020). Heterogeneous representations in the hippocampus.
Roy, D.S., Arons, A., Mitchell, T.I., Pignatelli, M., Ryan, T.J., and Tonegawa, S. Neurosci. Res. Published online May 20, 2020. https://doi.org/10.1016/j.neu-
(2016). Memory retrieval by activating engram cells in mouse models of early res.2020.05.002.
Alzheimer’s disease. Nature 531, 508–512.
Tanaka, K.Z., and McHugh, T.J. (2018). The hippocampal engram as a memory
Roy, D.S., Muralidhar, S., Smith, L.M., and Tonegawa, S. (2017). Silent mem- index. J. Exp. Neurosci. 12, 1179069518815942.
ory engrams as the basis for retrograde amnesia. Proc. Natl. Acad. Sci. USA
114, E9972–E9979. Tanaka, K.Z., Pevzner, A., Hamidi, A.B., Nakazawa, Y., Graham, J., and Wilt-
gen, B.J. (2014). Cortical representations are reinstated by the hippocampus
Roy, D.S., Park, Y.-G., Ogawa, S.K., Cho, J.H., Choi, H., Kamensky, L., Martin, during memory retrieval. Neuron 84, 347–354.
J., Chung, K., and Tonegawa, S. (2019). Brain-wide mapping of contextual fear
memory engram ensembles supports the dispersed engram complex hypoth- Tanaka, K.Z., He, H., Tomar, A., Niisato, K., Huang, A.J.Y., and McHugh, T.J.
esis. bioRxiv. https://doi.org/10.1101/668483. (2018). The hippocampal engram maps experience but not place. Science
361, 392–397.
Royer, S., Zemelman, B.V., Losonczy, A., Kim, J., Chance, F., Magee, J.C.,
and Buzsáki, G. (2012). Control of timing, rate and bursts of hippocampal place Tayler, K.K., Tanaka, K.Z., Reijmers, L.G., and Wiltgen, B.J. (2013). Reactiva-
cells by dendritic and somatic inhibition. Nat. Neurosci. 15, 769–775. tion of neural ensembles during the retrieval of recent and remote memory.
Curr. Biol. 23, 99–106.
Rudy, J.W., and O’Reilly, R.C. (1999). Contextual fear conditioning, conjunc-
tive representations, pattern completion, and the hippocampus. Behav. Neu- Teyler, T.J., and DiScenna, P. (1985). The role of hippocampus in memory: a
rosci. 113, 867–880. hypothesis. Neurosci. Biobehav. Rev. 9, 377–389.

Ryan, T.J., Roy, D.S., Pignatelli, M., Arons, A., and Tonegawa, S. (2015). Mem- Teyler, T.J., and DiScenna, P. (1986). The hippocampal memory indexing the-
ory. Engram cells retain memory under retrograde amnesia. Science 348, ory. Behav. Neurosci. 100, 147–154.
1007–1013.
Teyler, T.J., and Rudy, J.W. (2007). The hippocampal indexing theory and
Schacter, D.L., Eich, J.E., and Tulving, E. (1978). Richard Semon’s theory of episodic memory: updating the index. Hippocampus 17, 1158–1169.
memory. J. Verbal Learn. Verbal Behav. 17, 721–743.
Tingley, D., and Buzsáki, G. (2018). Transformation of a spatial map across the
hippocampal-lateral septal circuit. Neuron 98, 1229–1242.e5.
Schultz, H., Tibon, R., LaRocque, K.F., Gagnon, S.A., Wagner, A.D., and Star-
esina, B.P. (2019). Content tuning in the medial temporal lobe cortex: voxels Tonegawa, S., Morrissey, M.D., and Kitamura, T. (2018). The role of engram
that perceive, retrieve. eNeuro 6, ENEURO.0291-19.2019. cells in the systems consolidation of memory. Nat. Rev. Neurosci. 19,
485–498.
Semon, R.W. (1921). The Mneme (The Macmillan Company).
Trelle, A.N., Carr, V.A., Guerin, S.A., Thieu, M.K., Jayakumar, M., Guo, W., Na-
Simkin, D., Hattori, S., Ybarra, N., Musial, T.F., Buss, E.W., Richter, H., Oh, diadwala, A., Corso, N.K., Hunt, M.P., Litovsky, C.P., et al. (2020). Hippocam-
M.M., Nicholson, D.A., and Disterhoft, J.F. (2015). Aging-related hyperexcit- pal and cortical mechanisms at retrieval explain variability in episodic remem-
ability in CA3 pyramidal neurons is mediated by enhanced A-type K+ channel bering in older adults. eLife 9, 9.
function and expression. J. Neurosci. 35, 13206–13218.
Trouche, S., Perestenko, P.V., van de Ven, G.M., Bratley, C.T., McNamara,
Skaggs, W.E., and McNaughton, B.L. (1992). Computational approaches to C.G., Campo-Urriza, N., Black, S.L., Reijmers, L.G., and Dupret, D. (2016). Re-
hippocampal function. Curr. Opin. Neurobiol. 2, 209–211. coding a cocaine-place memory engram to a neutral engram in the hippocam-
pus. Nat. Neurosci. 19, 564–567.
Small, S.A., Schobel, S.A., Buxton, R.B., Witter, M.P., and Barnes, C.A. (2011).
A pathophysiological framework of hippocampal dysfunction in ageing and Trouche, S., Koren, V., Doig, N.M., Ellender, T.J., El-Gaby, M., Lopes-Dos-
disease. Nat. Rev. Neurosci. 12, 585–601. Santos, V., Reeve, H.M., Perestenko, P.V., Garas, F.N., Magill, P.J., et al.
(2019). A hippocampus-accumbens tripartite neuronal motif guides appetitive
Soltesz, I., and Losonczy, A. (2018). CA1 pyramidal cell diversity enabling par- memory in space. Cell 176, 1393–1406.e16.
allel information processing in the hippocampus. Nat. Neurosci. 21, 484–493.
Tulving, E. (2002). Episodic memory: from mind to brain. Annu. Rev. Psychol.
Sosa, M., Gillespie, A.K., and Frank, L.M. (2018). Neural activity patterns un- 53, 1–25.
derlying spatial coding in the hippocampus. Curr. Top. Behav. Neurosci.
37, 43–100. Tyssowski, K.M., DeStefino, N.R., Cho, J.-H., Dunn, C.J., Poston, R.G., Carty,
C.E., Jones, R.D., Chang, S.M., Romeo, P., Wurzelmann, M.K., et al. (2018).
Squire, L.R., Stark, C.E.L., and Clark, R.E. (2004). The medial temporal lobe. Different neuronal activity patterns induce different gene expression pro-
Annu. Rev. Neurosci. 27, 279–306. grams. Neuron 98, 530–546.e11.

Neuron 107, September 9, 2020 15


Please cite this article in press as: Goode et al., An Integrated Index: Engrams, Place Cells, and Hippocampal Memory, Neuron (2020), https://doi.org/
10.1016/j.neuron.2020.07.011

ll
Perspective
Valero, M., and de la Prida, L.M. (2018). The hippocampus in depth: a sublayer- Wirt, R.A., and Hyman, J.M. (2019). ACC theta improves hippocampal contex-
specific perspective of entorhinal-hippocampal function. Curr. Opin. Neuro- tual processing during remote recall. Cell Rep. 27, 2313–2327.e4.
biol. 52, 107–114.
Wixted, J.T., Squire, L.R., Jang, Y., Papesh, M.H., Goldinger, S.D., Kuhn, J.R.,
Viana da Silva, S., Zhang, P., Haberl, M.G., Labrousse, V., Grosjean, N., Blan- Smith, K.A., Treiman, D.M., and Steinmetz, P.N. (2014). Sparse and distributed
chet, C., Frick, A., and Mulle, C. (2019). Hippocampal mossy fibers synapses in coding of episodic memory in neurons of the human hippocampus. Proc. Natl.
CA3 pyramidal cells are altered at an early stage in a mouse model of alz- Acad. Sci. USA 111, 9621–9626.
heimer’s disease. J. Neurosci. 39, 4193–4205.

Vinogradova, O.S. (2001). Hippocampus as comparator: role of the two input Wixted, J.T., Goldinger, S.D., Squire, L.R., Kuhn, J.R., Papesh, M.H., Smith,
and two output systems of the hippocampus in selection and registration of in- K.A., Treiman, D.M., and Steinmetz, P.N. (2018). Coding of episodic memory
formation. Hippocampus 11, 578–598. in the human hippocampus. Proc. Natl. Acad. Sci. USA 115, 1093–1098.

Wagatsuma, A., Okuyama, T., Sun, C., Smith, L.M., Abe, K., and Tonegawa, S. Yonelinas, A.P., Ranganath, C., Ekstrom, A.D., and Wiltgen, B.J. (2019). A
(2018). Locus coeruleus input to hippocampal CA3 drives single-trial learning contextual binding theory of episodic memory: systems consolidation recon-
of a novel context. Proc. Natl. Acad. Sci. USA 115, E310–E316. sidered. Nat. Rev. Neurosci. 20, 364–375.

Wang, S.-H., Teixeira, C.M., Wheeler, A.L., and Frankland, P.W. (2009). The Zhang, T.R., Larosa, A., Di Raddo, M.-E., Wong, V., Wong, A.S., and Wong,
precision of remote context memories does not require the hippocampus.
T.P. (2019). Negative memory engrams in the hippocampus enhance the sus-
Nat. Neurosci. 12, 253–255. ceptibility to chronic social defeat stress. J. Neurosci. 39, 7576–7590.
Wikenheiser, A.M., and Redish, A.D. (2015). Decoding the cognitive map:
ensemble hippocampal sequences and decision making. Curr. Opin. Neuro- Zhou, W., He, Y., Rehman, A.U., Kong, Y., Hong, S., Ding, G., Yalamanchili,
biol. 32, 8–15. H.K., Wan, Y.-W., Paul, B., Wang, C., et al.; DDD study (2019a). Loss of func-
tion of NCOR1 and NCOR2 impairs memory through a novel GABAergic hypo-
Wilson, I.A., Ikonen, S., Gallagher, M., Eichenbaum, H., and Tanila, H. (2005). thalamus-CA3 projection. Nat. Neurosci. 22, 205–217.
Age-associated alterations of hippocampal place cells are subregion specific.
J. Neurosci. 25, 6877–6886. Zhou, Y., Zhu, H., Liu, Z., Chen, X., Su, X., Ma, C., Tian, Z., Huang, B., Yan, E.,
Liu, X., and Ma, L. (2019b). A ventral CA1 to nucleus accumbens core engram
Wiltgen, B.J., Sanders, M.J., Behne, N.S., and Fanselow, M.S. (2001). Sex dif- circuit mediates conditioned place preference for cocaine. Nat. Neurosci. 22,
ferences, context preexposure, and the immediate shock deficit in Pavlovian 1986–1999.
context conditioning with mice. Behav. Neurosci. 115, 26–32.

Wiltgen, B.J., Zhou, M., Cai, Y., Balaji, J., Karlsson, M.G., Parivash, S.N., Li, Zhu, X.O., McCabe, B.J., Aggleton, J.P., and Brown, M.W. (1997). Differential
W., and Silva, A.J. (2010). The hippocampus plays a selective role in the activation of the rat hippocampus and perirhinal cortex by novel visual stimuli
retrieval of detailed contextual memories. Curr. Biol. 20, 1336–1344. and a novel environment. Neurosci. Lett. 229, 141–143.

Winocur, G., Moscovitch, M., and Sekeres, M. (2007). Memory consolidation Ziv, Y., Burns, L.D., Cocker, E.D., Hamel, E.O., Ghosh, K.K., Kitch, L.J., El Ga-
or transformation: context manipulation and hippocampal representations of mal, A., and Schnitzer, M.J. (2013). Long-term dynamics of CA1 hippocampal
memory. Nat. Neurosci. 10, 555–557. place codes. Nat. Neurosci. 16, 264–266.

16 Neuron 107, September 9, 2020


Another random document with
no related content on Scribd:
The canal is deserted. The music-boats have long since put out their
lanterns and tied up for the night. The lighters at the Dogana
opposite lie still and motionless, their crews asleep under the mats
stretched on the decks. Away up in the blue swims the silver moon,
attended by an escort of clouds hovering close about her. Towering
above you rises the great dome of the Salute, silent, majestic, every
statue, cross, and scroll bathed in the glory of her light.
Suddenly, as you hang over your balcony, the soft night embracing
you, the odor of oleanders filling the air, you hear the quick
movement of a flute borne on the night wind from away up the Iron
Bridge. Nearer it comes, nearer, the clear, bird-like notes floating
over the still canal and the deserted city. You lean forward and catch
the spring and rhythm of the two gondoliers as they glide past,
keeping time to the thrill of the melody. You catch, too, the abandon
and charm of it all. He is standing over her, his head uncovered, the
moonlight glinting on the uplifted reed at his lips. She lies on the
cushions beneath him, throat and shoulders bare, a light scarf about
her head. It is only a glimpse, but it lingers in your memory for years,
—you on the balcony and alone.
Out they go,—out into the wide lagoon,—out into the soft night,
under the glory of the radiant stars. Fainter and fainter falls the
music, dimmer and dimmer pales the speck with its wake of silver.
Then all is still!
The Riverside Press
CAMBRIDGE, MASSACHUSETTS, U. S. A.
ELECTROTYPED AND PRINTED BY
H. O. HOUGHTON AND CO.
FOOTNOTE:
[1] Misteri di Venezia, di Edmondo Lundy.
TRANSCRIBER’S NOTES:
Obvious typographical errors have been corrected.
Inconsistencies in hyphenation have been
standardized.
Archaic or variant spelling has been retained.
*** END OF THE PROJECT GUTENBERG EBOOK GONDOLA
DAYS ***

Updated editions will replace the previous one—the old editions


will be renamed.

Creating the works from print editions not protected by U.S.


copyright law means that no one owns a United States copyright
in these works, so the Foundation (and you!) can copy and
distribute it in the United States without permission and without
paying copyright royalties. Special rules, set forth in the General
Terms of Use part of this license, apply to copying and
distributing Project Gutenberg™ electronic works to protect the
PROJECT GUTENBERG™ concept and trademark. Project
Gutenberg is a registered trademark, and may not be used if
you charge for an eBook, except by following the terms of the
trademark license, including paying royalties for use of the
Project Gutenberg trademark. If you do not charge anything for
copies of this eBook, complying with the trademark license is
very easy. You may use this eBook for nearly any purpose such
as creation of derivative works, reports, performances and
research. Project Gutenberg eBooks may be modified and
printed and given away—you may do practically ANYTHING in
the United States with eBooks not protected by U.S. copyright
law. Redistribution is subject to the trademark license, especially
commercial redistribution.

START: FULL LICENSE


THE FULL PROJECT GUTENBERG LICENSE
PLEASE READ THIS BEFORE YOU DISTRIBUTE OR USE THIS WORK

To protect the Project Gutenberg™ mission of promoting the


free distribution of electronic works, by using or distributing this
work (or any other work associated in any way with the phrase
“Project Gutenberg”), you agree to comply with all the terms of
the Full Project Gutenberg™ License available with this file or
online at www.gutenberg.org/license.

Section 1. General Terms of Use and


Redistributing Project Gutenberg™
electronic works
1.A. By reading or using any part of this Project Gutenberg™
electronic work, you indicate that you have read, understand,
agree to and accept all the terms of this license and intellectual
property (trademark/copyright) agreement. If you do not agree to
abide by all the terms of this agreement, you must cease using
and return or destroy all copies of Project Gutenberg™
electronic works in your possession. If you paid a fee for
obtaining a copy of or access to a Project Gutenberg™
electronic work and you do not agree to be bound by the terms
of this agreement, you may obtain a refund from the person or
entity to whom you paid the fee as set forth in paragraph 1.E.8.

1.B. “Project Gutenberg” is a registered trademark. It may only


be used on or associated in any way with an electronic work by
people who agree to be bound by the terms of this agreement.
There are a few things that you can do with most Project
Gutenberg™ electronic works even without complying with the
full terms of this agreement. See paragraph 1.C below. There
are a lot of things you can do with Project Gutenberg™
electronic works if you follow the terms of this agreement and
help preserve free future access to Project Gutenberg™
electronic works. See paragraph 1.E below.
1.C. The Project Gutenberg Literary Archive Foundation (“the
Foundation” or PGLAF), owns a compilation copyright in the
collection of Project Gutenberg™ electronic works. Nearly all the
individual works in the collection are in the public domain in the
United States. If an individual work is unprotected by copyright
law in the United States and you are located in the United
States, we do not claim a right to prevent you from copying,
distributing, performing, displaying or creating derivative works
based on the work as long as all references to Project
Gutenberg are removed. Of course, we hope that you will
support the Project Gutenberg™ mission of promoting free
access to electronic works by freely sharing Project
Gutenberg™ works in compliance with the terms of this
agreement for keeping the Project Gutenberg™ name
associated with the work. You can easily comply with the terms
of this agreement by keeping this work in the same format with
its attached full Project Gutenberg™ License when you share it
without charge with others.

1.D. The copyright laws of the place where you are located also
govern what you can do with this work. Copyright laws in most
countries are in a constant state of change. If you are outside
the United States, check the laws of your country in addition to
the terms of this agreement before downloading, copying,
displaying, performing, distributing or creating derivative works
based on this work or any other Project Gutenberg™ work. The
Foundation makes no representations concerning the copyright
status of any work in any country other than the United States.

1.E. Unless you have removed all references to Project


Gutenberg:

1.E.1. The following sentence, with active links to, or other


immediate access to, the full Project Gutenberg™ License must
appear prominently whenever any copy of a Project
Gutenberg™ work (any work on which the phrase “Project
Gutenberg” appears, or with which the phrase “Project
Gutenberg” is associated) is accessed, displayed, performed,
viewed, copied or distributed:

This eBook is for the use of anyone anywhere in the United


States and most other parts of the world at no cost and with
almost no restrictions whatsoever. You may copy it, give it
away or re-use it under the terms of the Project Gutenberg
License included with this eBook or online at
www.gutenberg.org. If you are not located in the United
States, you will have to check the laws of the country where
you are located before using this eBook.

1.E.2. If an individual Project Gutenberg™ electronic work is


derived from texts not protected by U.S. copyright law (does not
contain a notice indicating that it is posted with permission of the
copyright holder), the work can be copied and distributed to
anyone in the United States without paying any fees or charges.
If you are redistributing or providing access to a work with the
phrase “Project Gutenberg” associated with or appearing on the
work, you must comply either with the requirements of
paragraphs 1.E.1 through 1.E.7 or obtain permission for the use
of the work and the Project Gutenberg™ trademark as set forth
in paragraphs 1.E.8 or 1.E.9.

1.E.3. If an individual Project Gutenberg™ electronic work is


posted with the permission of the copyright holder, your use and
distribution must comply with both paragraphs 1.E.1 through
1.E.7 and any additional terms imposed by the copyright holder.
Additional terms will be linked to the Project Gutenberg™
License for all works posted with the permission of the copyright
holder found at the beginning of this work.

1.E.4. Do not unlink or detach or remove the full Project


Gutenberg™ License terms from this work, or any files
containing a part of this work or any other work associated with
Project Gutenberg™.
1.E.5. Do not copy, display, perform, distribute or redistribute
this electronic work, or any part of this electronic work, without
prominently displaying the sentence set forth in paragraph 1.E.1
with active links or immediate access to the full terms of the
Project Gutenberg™ License.

1.E.6. You may convert to and distribute this work in any binary,
compressed, marked up, nonproprietary or proprietary form,
including any word processing or hypertext form. However, if
you provide access to or distribute copies of a Project
Gutenberg™ work in a format other than “Plain Vanilla ASCII” or
other format used in the official version posted on the official
Project Gutenberg™ website (www.gutenberg.org), you must, at
no additional cost, fee or expense to the user, provide a copy, a
means of exporting a copy, or a means of obtaining a copy upon
request, of the work in its original “Plain Vanilla ASCII” or other
form. Any alternate format must include the full Project
Gutenberg™ License as specified in paragraph 1.E.1.

1.E.7. Do not charge a fee for access to, viewing, displaying,


performing, copying or distributing any Project Gutenberg™
works unless you comply with paragraph 1.E.8 or 1.E.9.

1.E.8. You may charge a reasonable fee for copies of or


providing access to or distributing Project Gutenberg™
electronic works provided that:

• You pay a royalty fee of 20% of the gross profits you derive from
the use of Project Gutenberg™ works calculated using the
method you already use to calculate your applicable taxes. The
fee is owed to the owner of the Project Gutenberg™ trademark,
but he has agreed to donate royalties under this paragraph to
the Project Gutenberg Literary Archive Foundation. Royalty
payments must be paid within 60 days following each date on
which you prepare (or are legally required to prepare) your
periodic tax returns. Royalty payments should be clearly marked
as such and sent to the Project Gutenberg Literary Archive
Foundation at the address specified in Section 4, “Information
about donations to the Project Gutenberg Literary Archive
Foundation.”

• You provide a full refund of any money paid by a user who


notifies you in writing (or by e-mail) within 30 days of receipt that
s/he does not agree to the terms of the full Project Gutenberg™
License. You must require such a user to return or destroy all
copies of the works possessed in a physical medium and
discontinue all use of and all access to other copies of Project
Gutenberg™ works.

• You provide, in accordance with paragraph 1.F.3, a full refund of


any money paid for a work or a replacement copy, if a defect in
the electronic work is discovered and reported to you within 90
days of receipt of the work.

• You comply with all other terms of this agreement for free
distribution of Project Gutenberg™ works.

1.E.9. If you wish to charge a fee or distribute a Project


Gutenberg™ electronic work or group of works on different
terms than are set forth in this agreement, you must obtain
permission in writing from the Project Gutenberg Literary
Archive Foundation, the manager of the Project Gutenberg™
trademark. Contact the Foundation as set forth in Section 3
below.

1.F.

1.F.1. Project Gutenberg volunteers and employees expend


considerable effort to identify, do copyright research on,
transcribe and proofread works not protected by U.S. copyright
law in creating the Project Gutenberg™ collection. Despite
these efforts, Project Gutenberg™ electronic works, and the
medium on which they may be stored, may contain “Defects,”
such as, but not limited to, incomplete, inaccurate or corrupt
data, transcription errors, a copyright or other intellectual
property infringement, a defective or damaged disk or other
medium, a computer virus, or computer codes that damage or
cannot be read by your equipment.

1.F.2. LIMITED WARRANTY, DISCLAIMER OF DAMAGES -


Except for the “Right of Replacement or Refund” described in
paragraph 1.F.3, the Project Gutenberg Literary Archive
Foundation, the owner of the Project Gutenberg™ trademark,
and any other party distributing a Project Gutenberg™ electronic
work under this agreement, disclaim all liability to you for
damages, costs and expenses, including legal fees. YOU
AGREE THAT YOU HAVE NO REMEDIES FOR NEGLIGENCE,
STRICT LIABILITY, BREACH OF WARRANTY OR BREACH
OF CONTRACT EXCEPT THOSE PROVIDED IN PARAGRAPH
1.F.3. YOU AGREE THAT THE FOUNDATION, THE
TRADEMARK OWNER, AND ANY DISTRIBUTOR UNDER
THIS AGREEMENT WILL NOT BE LIABLE TO YOU FOR
ACTUAL, DIRECT, INDIRECT, CONSEQUENTIAL, PUNITIVE
OR INCIDENTAL DAMAGES EVEN IF YOU GIVE NOTICE OF
THE POSSIBILITY OF SUCH DAMAGE.

1.F.3. LIMITED RIGHT OF REPLACEMENT OR REFUND - If


you discover a defect in this electronic work within 90 days of
receiving it, you can receive a refund of the money (if any) you
paid for it by sending a written explanation to the person you
received the work from. If you received the work on a physical
medium, you must return the medium with your written
explanation. The person or entity that provided you with the
defective work may elect to provide a replacement copy in lieu
of a refund. If you received the work electronically, the person or
entity providing it to you may choose to give you a second
opportunity to receive the work electronically in lieu of a refund.
If the second copy is also defective, you may demand a refund
in writing without further opportunities to fix the problem.

1.F.4. Except for the limited right of replacement or refund set


forth in paragraph 1.F.3, this work is provided to you ‘AS-IS’,
WITH NO OTHER WARRANTIES OF ANY KIND, EXPRESS
OR IMPLIED, INCLUDING BUT NOT LIMITED TO
WARRANTIES OF MERCHANTABILITY OR FITNESS FOR
ANY PURPOSE.

1.F.5. Some states do not allow disclaimers of certain implied


warranties or the exclusion or limitation of certain types of
damages. If any disclaimer or limitation set forth in this
agreement violates the law of the state applicable to this
agreement, the agreement shall be interpreted to make the
maximum disclaimer or limitation permitted by the applicable
state law. The invalidity or unenforceability of any provision of
this agreement shall not void the remaining provisions.

1.F.6. INDEMNITY - You agree to indemnify and hold the


Foundation, the trademark owner, any agent or employee of the
Foundation, anyone providing copies of Project Gutenberg™
electronic works in accordance with this agreement, and any
volunteers associated with the production, promotion and
distribution of Project Gutenberg™ electronic works, harmless
from all liability, costs and expenses, including legal fees, that
arise directly or indirectly from any of the following which you do
or cause to occur: (a) distribution of this or any Project
Gutenberg™ work, (b) alteration, modification, or additions or
deletions to any Project Gutenberg™ work, and (c) any Defect
you cause.

Section 2. Information about the Mission of


Project Gutenberg™
Project Gutenberg™ is synonymous with the free distribution of
electronic works in formats readable by the widest variety of
computers including obsolete, old, middle-aged and new
computers. It exists because of the efforts of hundreds of
volunteers and donations from people in all walks of life.

Volunteers and financial support to provide volunteers with the


assistance they need are critical to reaching Project
Gutenberg™’s goals and ensuring that the Project Gutenberg™
collection will remain freely available for generations to come. In
2001, the Project Gutenberg Literary Archive Foundation was
created to provide a secure and permanent future for Project
Gutenberg™ and future generations. To learn more about the
Project Gutenberg Literary Archive Foundation and how your
efforts and donations can help, see Sections 3 and 4 and the
Foundation information page at www.gutenberg.org.

Section 3. Information about the Project


Gutenberg Literary Archive Foundation
The Project Gutenberg Literary Archive Foundation is a non-
profit 501(c)(3) educational corporation organized under the
laws of the state of Mississippi and granted tax exempt status by
the Internal Revenue Service. The Foundation’s EIN or federal
tax identification number is 64-6221541. Contributions to the
Project Gutenberg Literary Archive Foundation are tax
deductible to the full extent permitted by U.S. federal laws and
your state’s laws.

The Foundation’s business office is located at 809 North 1500


West, Salt Lake City, UT 84116, (801) 596-1887. Email contact
links and up to date contact information can be found at the
Foundation’s website and official page at
www.gutenberg.org/contact

Section 4. Information about Donations to


the Project Gutenberg Literary Archive
Foundation
Project Gutenberg™ depends upon and cannot survive without
widespread public support and donations to carry out its mission
of increasing the number of public domain and licensed works
that can be freely distributed in machine-readable form
accessible by the widest array of equipment including outdated
equipment. Many small donations ($1 to $5,000) are particularly
important to maintaining tax exempt status with the IRS.

The Foundation is committed to complying with the laws


regulating charities and charitable donations in all 50 states of
the United States. Compliance requirements are not uniform
and it takes a considerable effort, much paperwork and many
fees to meet and keep up with these requirements. We do not
solicit donations in locations where we have not received written
confirmation of compliance. To SEND DONATIONS or
determine the status of compliance for any particular state visit
www.gutenberg.org/donate.

While we cannot and do not solicit contributions from states


where we have not met the solicitation requirements, we know
of no prohibition against accepting unsolicited donations from
donors in such states who approach us with offers to donate.

International donations are gratefully accepted, but we cannot


make any statements concerning tax treatment of donations
received from outside the United States. U.S. laws alone swamp
our small staff.

Please check the Project Gutenberg web pages for current


donation methods and addresses. Donations are accepted in a
number of other ways including checks, online payments and
credit card donations. To donate, please visit:
www.gutenberg.org/donate.

Section 5. General Information About Project


Gutenberg™ electronic works
Professor Michael S. Hart was the originator of the Project
Gutenberg™ concept of a library of electronic works that could
be freely shared with anyone. For forty years, he produced and
distributed Project Gutenberg™ eBooks with only a loose
network of volunteer support.

Project Gutenberg™ eBooks are often created from several


printed editions, all of which are confirmed as not protected by
copyright in the U.S. unless a copyright notice is included. Thus,
we do not necessarily keep eBooks in compliance with any
particular paper edition.

Most people start at our website which has the main PG search
facility: www.gutenberg.org.

This website includes information about Project Gutenberg™,


including how to make donations to the Project Gutenberg
Literary Archive Foundation, how to help produce our new
eBooks, and how to subscribe to our email newsletter to hear
about new eBooks.

You might also like